Modern Trends in Structural and Solid Mechanics 1: Statics and Stability [1 ed.] 1786307146, 9781786307149

This book - comprised of three separate volumes - presents the recent developments and research discoveries in structura

221 101 18MB

English Pages 302 [291] Year 2021

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Modern Trends in Structural and Solid Mechanics 1: Statics and Stability [1 ed.]
 1786307146, 9781786307149

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Modern Trends in Structural and Solid Mechanics 1

Series Editor Noël Challamel

Modern Trends in Structural and Solid Mechanics 1 Statics and Stability

Edited by

Noël Challamel Julius Kaplunov Izuru Takewaki

First published 2021 in Great Britain and the United States by ISTE Ltd and John Wiley & Sons, Inc.

Apart from any fair dealing for the purposes of research or private study, or criticism or review, as permitted under the Copyright, Designs and Patents Act 1988, this publication may only be reproduced, stored or transmitted, in any form or by any means, with the prior permission in writing of the publishers, or in the case of reprographic reproduction in accordance with the terms and licenses issued by the CLA. Enquiries concerning reproduction outside these terms should be sent to the publishers at the undermentioned address: ISTE Ltd 27-37 St George’s Road London SW19 4EU UK

John Wiley & Sons, Inc. 111 River Street Hoboken, NJ 07030 USA

www.iste.co.uk

www.wiley.com

© ISTE Ltd 2021 The rights of Noël Challamel, Julius Kaplunov and Izuru Takewaki to be identified as the authors of this work have been asserted by them in accordance with the Copyright, Designs and Patents Act 1988. Library of Congress Control Number: 2020952688 British Library Cataloguing-in-Publication Data A CIP record for this book is available from the British Library ISBN 978-1-78630-714-9

Contents

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

xi

Noël CHALLAMEL, Julius KAPLUNOV and Izuru TAKEWAKI Chapter 1. Static Deformations of Fiber-Reinforced Composite Laminates by the Least-Squares Method . . . . . . . . . . . . . . . . . . . . . . .

1

Devin BURNS and Romesh C. BATRA 1.1. Introduction . . . . . . . . . . . . . . . . . . . . . 1.2. Formulation of the problem . . . . . . . . . . . . . 1.3. Results and discussion . . . . . . . . . . . . . . . . 1.3.1. Verification of the numerical algorithm. . . . 1.3.2. Simply supported sandwich plate . . . . . . . 1.3.3. Laminate with arbitrary boundary conditions 1.4. Remarks . . . . . . . . . . . . . . . . . . . . . . . 1.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . 1.6. Acknowledgments . . . . . . . . . . . . . . . . . . 1.7. References . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

Chapter 2. Stability of Laterally Compressed Elastic Chains . . . . . . . . .

2 3 7 7 9 10 14 15 15 15 17

Andrii I AKOVLIEV , Srinandan DASMAHAPATRA and Atul B HASKAR 2.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Compression of stacked elastic sheets . . . . . . . . . . . . 2.3. Stability of an elastically coupled cyclic chain . . . . . . . 2.4. Elastic stability of two coupled rods with disorder . . . . . 2.5. Spatial localization of lateral buckling in a disordered coupled rigid rods . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.7. References . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . chain . . . . . . . . . . . .

. . . . . . . . of . . . . . .

. . . . . . . . . . . . . . . . . . . . elastically . . . . . . . . . . . . . . .

17 20 23 28 31 39 40

vi

Modern Trends in Structural and Solid Mechanics 1

Chapter 3. Analysis of a Beck’s Column over Fractional-Order Restraints via Extended Routh–Hurwitz Theorem . . . . . . . . . . . . . . . . . . . . . . .

43

Emanuela BOLOGNA, Mario DI PAOLA, Massimiliano ZINGALES 3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Material hereditariness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1. Linear hereditariness: fractional-order models . . . . . . . . . . . . . . . . 3.3. Dynamic equilibrium of an elastic cantilever over a fractional-order foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4. Stability analysis of Beck’s column over fractional-order hereditary foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1. The characteristic polynomial . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.2. State-space representation of the dynamic equilibrium equation . . . . . . 3.4.3. Stability analysis of fractional-order Beck’s column via the extended Routh–Hurwitz criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5. Numerical application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 4. Localization in the Static Response of Higher-Order Lattices with Long-Range Interactions . . . . . . . . . . . . . . . . . . . . . . . . .

43 44 48 51 54 55 57 60 63 65 65

67

Noël CHALLAMEL and Vincent PICANDET 4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Two-neighbor interaction – general formulation – homogeneous solution . 4.3. Two-neighbor interaction – localization in a weakened problem . . . . . . 4.4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

Chapter 5. New Analytic Solutions for Elastic Buckling of Isotropic Plates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

68 70 76 86 86

91

Joseph T ENENBAUM , Aharon D EUTSCH and Moshe E ISENBERGER 5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Equilibrium equation . . . . . . . . . . . . . . . . . . . . . . . 5.3. Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4. Boundary condition . . . . . . . . . . . . . . . . . . . . . . . . 5.5. Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . 5.6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7. Appendix A: Deflection, slopes, bending moments and shears . 5.8. Appendix B: Function transformation . . . . . . . . . . . . . . 5.9. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

91 93 94 96 98 105 109 116 119

Contents

Chapter 6. Buckling and Post-Buckling of Parabolic Arches with Local Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

vii

121

Uğurcan EROĞLU, Giuseppe RUTA, Achille PAOLONE and Ekrem T ÜFEKCI 6.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . 6.2. A one-dimensional model for arches . . . . . . . . . . 6.2.1. Finite kinematics and balance, linear elastic law . 6.2.2. Non-trivial fundamental equilibrium path . . . . . 6.2.3. Bifurcated path . . . . . . . . . . . . . . . . . . . 6.2.4. Special benchmark examples. . . . . . . . . . . . 6.3. Parabolic arches . . . . . . . . . . . . . . . . . . . . . 6.4. Crack models for one-dimensional elements . . . . . 6.5. An application . . . . . . . . . . . . . . . . . . . . . . 6.5.1. A comparison . . . . . . . . . . . . . . . . . . . . 6.6. Final remarks . . . . . . . . . . . . . . . . . . . . . . . 6.7. Acknowledgments . . . . . . . . . . . . . . . . . . . . 6.8. References . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

Chapter 7. Inelastic Microbuckling of Composites by Wave-Buckling Analogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

121 122 124 126 127 128 130 133 135 138 138 139 139

145

Rivka GILAT and Jacob A BOUDI 7.1. Introduction . . . . . . . . . . . . . . . . . . . . 7.2. Buckling-wave propagation analogy . . . . . . . 7.3. Microbuckling in elastic orthotropic composites 7.4. Inelastic microbuckling . . . . . . . . . . . . . . 7.5. Results and discussion . . . . . . . . . . . . . . . 7.6. References . . . . . . . . . . . . . . . . . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

Chapter 8. Quasi-Bifurcation of Discrete Systems with Unstable Post-Critical Behavior under Impulsive Loads . . . . . . . . . . . . . . . . . . . .

145 146 148 150 152 156

159

Mariano P. AMEIJEIRAS and Luis A. GODOY 8.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2. Case study of a two DOF system with unstable static behavior . . . 8.3. Exploring the static and dynamic behavior of the two DOF system. 8.4. The dynamic stability criterion due to Lee . . . . . . . . . . . . . . 8.5. New stability bounds following Lee’s approach . . . . . . . . . . . 8.6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7. Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.8. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

160 161 164 167 170 174 175 175

viii

Modern Trends in Structural and Solid Mechanics 1

Chapter 9. Singularly Perturbed Problems of Drill String Buckling in Deep Curvilinear Borehole Channels . . . . . . . . . . . . . . . . . . . . . . . .

177

Valery I. GULYAYEV and Natalya V. SHLYUN 9.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2. Singular perturbation theory: elements and history . . . . . . . . . . . . . 9.3. Posing the problem of a drill string buckling in the curvilinear borehole . 9.4. Modeling the drill string buckling in lowering operation . . . . . . . . . 9.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . .

Chapter 10. Shape-optimized Cantilevered Columns under a Rocket-based Follower Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

177 179 184 195 199

201

Yoshihiko SUGIYAMA, Mikael A. LANGTHJEM and Kazuo KATAYAMA 10.1. Background . . . . . . . 10.2. Aims . . . . . . . . . . . 10.3. Numerical analysis . . . 10.3.1. Stability analysis . . 10.3.2. Optimum design . . 10.4. Experiment . . . . . . . 10.4.1. General description 10.4.2. Rocket motor . . . . 10.4.3. Columns . . . . . . 10.4.4. Free vibration test . 10.5. Flutter test . . . . . . . . 10.6. Concluding remarks . . 10.7. Acknowledgments . . . 10.8. Appendix . . . . . . . . 10.9. References . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

Chapter 11. Hencky Bar-Chain Model for Buckling Analysis and Optimal Design of Trapezoidal Arches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

201 204 206 206 209 213 213 214 216 219 220 223 224 224 225

229

Chien Ming WANG, Wen Hao PAN and Hanzhe ZHANG 11.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2. Buckling analysis of trapezoidal arches based on the HBM . . 11.2.1. Description of the HBM . . . . . . . . . . . . . . . . . . . 11.2.2. HBM stiffness matrix formulation . . . . . . . . . . . . . . 11.2.3. Governing equation considering compatibility conditions . 11.2.4. Verification of the HBM . . . . . . . . . . . . . . . . . . . 11.3. Optimal design of symmetric trapezoidal arches . . . . . . . . 11.3.1. Problem definition . . . . . . . . . . . . . . . . . . . . . . . 11.3.2. Optimization procedure . . . . . . . . . . . . . . . . . . . . 11.3.3. Optimal solutions . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

230 231 232 234 235 236 239 239 240 240

Contents

11.3.4. Sensitivity analysis of optimal solutions. . . . . . . . . . . . 11.3.5. Comparison with the buckling load of optimal fully stressed trapezoidal arches . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . 11.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

ix

. . . . . . .

243

. . . . . . . . . . . . . . . . . . . . .

245 245 246

List of Authors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

249

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

251

Summaries of Volumes 2 and 3 . . . . . . . . . . . . . . . . . . . . . . . . . . .

255

Preface Short Bibliographical Presentation of Prof. Isaac Elishakoff

This book is dedicated to Prof. Isaac Elishakoff by his colleagues, friends and former students, on the occasion of his seventy-fifth birthday.

Figure P.1. Prof. Isaac Elishakoff

For a color version of all the figures in this chapter, see www.iste.co.uk/challamel/ mechanics1.zip.

xii

Modern Trends in Structural and Solid Mechanics 1

Prof. Isaac Elishakoff is an international leading authority across a broad area of structural mechanics, including dynamics and stability, optimization and antioptimization, probabilistic methods, analysis of structures with uncertainty, refined theories, functionally graded material structures, and nanostructures. He was born in Kutaisi, Republic of Georgia, on February 9, 1944.

Figure P.2. Elishakoff in middle school in the city of Sukhumi, Georgia

Elishakoff holds a PhD in Dynamics and Strength of Machines from the Power Engineering Institute and Technical University in Moscow, Russia (Figure P.4 depicts the PhD defense of Prof. Isaac Elishakoff).

Figure P.3. Elishakoff just before acceptance to university. Photo taken in Sukhumi, Georgia

Preface

xiii

Figure P.4. Public PhD defense, Moscow Power Engineering Institute and State University; topic “Vibrational and Acoustical Fields in the Circular Cylindrical Shells Excited by Random Loadings”, and dedicated to the evaluation of noise levels in TU-144 supersonic aircraft

His supervisor was Prof. V. V. Bolotin (1926–2008), a member of the Russian Academy of Sciences (Figure P.5 shows Elishakoff with Bolotin some years later).

Figure P.5. Elishakoff with Bolotin (middle), member of the Russian Academy of Sciences, and Prof. Yukweng (Mike) Lin (left), member of the US National Academy of Engineering. Photo taken at Florida Atlantic University during a visit from Bolotin

Currently, Elishakoff is a Distinguished Research Professor in the Department of Ocean and Mechanical Engineering at Florida Atlantic University. Before joining the university, he taught for one year at Abkhazian University, Sukhumi in the

xiv

Modern Trends in Structural and Solid Mechanics 1

Republic of Georgia, and 18 years at the Technion – Israel Institute of Technology in Haifa, where he became the youngest full professor at the time of his promotion (Figure P.6 shows Elishakoff presenting a book to Prof. Josef Singer, Technion’s former president).

Figure P.6. Prof. Elishakoff presenting a book to Prof. J. Singer, Technion’s President; right: Prof. A. Libai, Aerospace Engineering Department, Technion

Elishakoff has lectured at about 200 meetings and seminars, including about 60 invited, plenary or keynote lectures, across Europe, North and South America, the Middle East and the Far East. Prof. Elishakoff has made vital and outstanding contributions in a number of areas in structural mechanics. In particular, he has analyzed random vibrations of homogeneous and composite beams, plates and shells, with special emphasis on the effects of refinements in structural theories and cross-correlations. Free structural vibrations have been tackled using a non-trivial generalization of Bolotin’s dynamic edge effect method. Nonlinear buckling has been investigated using a novel method, incorporating experimental analysis of imperfections. As a result, the fundamental concept of closing the gap – spanning the entire 20th century – between theory and practice in imperfection-sensitive structures has been proposed. Novel methods of evaluating structural reliability have been proposed, taking into account the error associated with various low-order approximations, as well as human error; innovative generalization of the stochastic linearization method has been advanced.

Preface

xv

A non-probabilistic theory for treating uncertainty in structural mechanics has been established. Dynamic stability of elastic and viscoelastic structures with imperfections has been studied. An improved, non-perturbative stochastic finite element method for structures has been developed. The list of Elishakoff’s remarkable research achievements goes on. His research has been acknowledged by many awards and prizes. He is a member of the European Academy of Sciences and Arts, a Fellow of the American Academy of Mechanics and ASME, and a Foreign Member of the Georgian National Academy of Sciences. Elishakoff is also a recipient of the Bathsheva de Rothschild prize (1973) and the Worcester Reed Warner Medal of the American Society of Mechanical Engineers (2016).

Figure P.7. Elishakoff having received the William B. Johnson Inter- Professional Founders Award

Elishakoff is directly involved in numerous editorial activities. He serves as the book review editor of the “Journal of Shock and Vibration” and is currently, or has previously been an associate editor of the International Journal of Mechanics of Machines and Structures, Applied Mechanics Reviews, and Chaos, Solitons & Fractals. In addition, he is or has been on the editorial boards of numerous journals, for

xvi

Modern Trends in Structural and Solid Mechanics 1

example Journal of Sound and Vibration, International Journal of Structural Stability and Dynamics, International Applied Mechanics and Computers & Structures. He also acts as a book series editor for Elsevier, Springer and Wiley.

Figure P.8. Inauguration as the Frank Freimann Visiting Professor of Aerospace and Mechanical Engineering; left: Rev. Theodore M. Hesburgh, President of the University of Notre Dame; right: Prof. Timothy O’Meara, Provost

Prof. Elishakoff has held prestigious visiting positions at top universities all over the world. Among them are Stanford University (S. P. Timoshenko Scholar); University of Notre Dame, USA (Frank M. Freimann Chair Professorship of Aerospace and Mechanical Engineering and Henry J. Massman, Jr. Chair Professorship of Civil Engineering); University of Palermo, Italy (Visiting Castigliano Distinguished Professor); Delft University of Technology, Netherlands (multiple appointments, including the W. T. Koiter Chair Professorship of the Mechanical Engineering Department – see Figure P.9); Universities of Tokyo and Kyoto, Japan (Fellow of the Japan Society for the Promotion of Science); Beijing University of Aeronautics and Astronautics, People’s Republic of China (Visiting Eminent Scholar); Technion, Haifa, Israel (Visiting Distinguished Professor); University of Southampton, UK (Distinguished Visiting Fellow of the Royal Academy of Engineering).

Preface

xvii

Figure P.9. Prof. Elishakoff with Prof. Warner Tjardus Koiter, Delft University of Technology (center), and Dr. V. Grishchak, of Ukraine (right)

Figure P.10. Elishakoff and his colleagues during the AIAA SDM Conference at Palm Springs, California in 2004; Standing, from right to left, are Prof. Elishakoff, the late Prof. Josef Singer and Dr. Giora Maymon of RAFAEL. Sitting is the late Prof. Avinoam Libai

xviii

Modern Trends in Structural and Solid Mechanics 1

Elishakoff has made a substantial contribution to conference organization. In particular, he participated in the organization of the Euro-Mech Colloquium on “Refined Dynamical Theories of Beams, Plates and Shells, and Their Applications” in Kassel, Germany (1986); the Second International Conference on Stochastic Structural Dynamics, in Boca Raton, USA (1990); “International Conference on Uncertain Structures” in Miami, USA and Western Caribbean (1996). He also coordinated four special courses at the International Centre for Mechanical Sciences (CISM), in Udine, Italy (1997, 2001, 2005, 2011). Prof. Elishakoff has published over 540 original papers in leading journals and conference proceedings. He championed authoring, co-authoring or editing of 31 influential and extremely well-received books and edited volumes. Here follows some praise of his work and books: – “It was not until 1979, when Elishakoff published his reliability study … that a method has been proposed, which made it possible to introduce the results of imperfection surveys … into the analysis …” (Prof. Johann Arbocz, Delft University of Technology, The Netherlands, Zeitschrift für Flugwissenschaften und Weltraumforschung). – “He has achieved world renown … His research is characterized by its originality and a combination of mathematical maturity and physical understanding which is reminiscent of von Kármán …” (Prof. Charles W. Bert, University of Oklahoma). – “It is clear that Elishakoff is a world leader in his field … His outstanding reputation is very well deserved …” (Prof. Bernard Budiansky, Harvard University). – “Professor Isaac Elishakoff … is subject-wise very much an all-round vibrationalist” (P. E. Doak, Editor in Chief, Journal of Sound and Vibration, University of Southampton, UK). – “This is a beautiful book …” (Dr. Stephen H. Crandall, Ford Professor of Engineering, M.I.T.). – “Das Buch ist in seiner Aufmachunghervorragendgestaltet und kannalsäusserstwertvolleErganzung … wäzmstensempfohlenwerden …” [The book’s appearance is perfectly designed and can be highly recommended as a valuable addition.] (Prof. Horst Försching, Institute of Aeroelasticity, Federal Republic of Germany, Zeitschrift für Flugwissenschaften und Weltraumforschung).

Preface

xix

– “Because of you, Notre Dame is an even better place, a more distinguished University” (Prof. Rev. Theodore M. Hesburgh, President, University of Notre Dame). – “It is an impressive volume …” (Prof. Warner T. Koiter, Delft University of Technology, The Netherlands). – “This extremely well-written text, authored by one of the leaders in the field, incorporates many of these new applications … Professor Elishakoff’s techniques for developing the material are accomplished in a way that illustrates his deep insight into the topic as well as his expertise as an educator … Clearly, the second half of the text provides the basis for an excellent graduate course in random vibrations and buckling … Professor Elishakoff has presented us with an outstanding instrument for teaching” (Prof. Frank Kozin, Polytechnic Institute of New York, American Institute of Aeronautics and Astronautics Journal). – “By far the best book on the market today …” (Prof. Niels C. Lind, University of Waterloo, Canada). – “The book develops a novel idea … Elegant, exhaustive discussion … The study can be an inspiration for further research, and provides excellent applications in design …” (Prof. G. A. Nariboli, Applied Mechanics Reviews). – “This volume is regarded as an advanced encyclopedia on random vibration and serves aeronautical, civil and mechanical engineers …” (Prof. Rauf Ibrahim, Wayne State University, Shock and Vibration Digest). – “The book deals with a fundamental problem in Applied Mechanics and in Engineering Sciences: How the uncertainties of the data of a problem influence its solution. The authors follow a novel approach for the treatment of these problems … The book is written with clarity and contains original and important results for the engineering sciences …” (Prof. P. D. Panagiotopoulos, University of Thessaloniki, Greece and University of Aachen, Germany, SIAM Review). – “The content should be of great interest to all engineers involved with vibration problems, placing the book well and truly in the category of an essential reference book …” (Prof. I. Pole, Journal of the British Society for Strain Measurement). – “A good book; a different book … It is hoped that the success of this book will encourage the author to provide a sequel in due course …” (Prof. John D. Robson, University of Glasgow, Scotland, UK, Journal of Sound and Vibration).

xx

Modern Trends in Structural and Solid Mechanics 1

– “The book certainly satisfies the need that now exists for a readable textbook and reference book …” (Prof. Masanobu Shinozuka, Columbia University). – “[the] author ties together reliability, random vibration and random buckling … Well written … useful book …” (Dr. H. Saunders, Shock and Vibration Digest). – “A very useful text that includes a broad spectrum of theory and application” (Mechanical Vibration, Prof. Haym Benaroya, Rutgers University). – “A treatise on random vibration and buckling … The reviewer wishes to compliment the author for the completion of a difficult task in preparing this book on a subject matter, which is still developing on many fronts …” (Prof. James T. P. Yao, Texas A&M University, Journal of Applied Mechanics). – “It seems to me a hard work with great result …” (Prof. Hans G. Natke, University of Hannover, Federal Republic of Germany). – “The approach is novel and could dominate the future practice of engineering” (The Structural Engineer). – “An excellent presentation … well written … all readers, students, and certainly reviewers should read this preface for its excellent presentation of the philosophy and raison d’être for this book. It is well written, with the material presented in an informational fashion as well as to raise questions related to unresolved … challenges; in the vernacular of film critics, ‘thumbs up’” (Dr. R. L. Sierakowski, U.S. Air Force Research Laboratory, AIAA Journal). – “This substantial and attractive volume is a well-organized and superbly written one that should be warmly welcomed by both theorists and practitioners … Prof. Elishakoff, Li, and Starnes, Jr. have given us a jewel of a book, one done with care and understanding of a complex and essential subject and one that seems to have ably filled a gap existing in the present-day literature and practice” (Current Engineering Practice). – “Most of the subjects covered in this outstanding book have never been discussed exclusively in the existing treatises … (Ocean Engineering). – “The treatment is scholarly, having about 900 items in the bibliography and additional contributors in the writing of almost every chapter … This reviewer believes that Non-Classical Problems in the Theory of Elastic Stability should be a useful reference for researchers, engineers, and graduate students in aeronautical,

Preface

xxi

mechanical, civil, nuclear, and marine engineering, and in applied mechanics” (Applied Mechanics Reviews). – “What more can be said about this monumental work, other than to express admiration? … The study is of great academic interest, and is clearly a labor of love. The author is to be congratulated on this work …” (Prof. H. D. Conway, Department of Theoretical and Applied Mechanics, Cornell University). – “This book … is prepared by Isaac Elishakoff, one of the eminent solid mechanics experts of the 20th century and the present one, and his distinguished coauthors, will be of enormous use to researchers, graduate students and professionals in the fields of ocean, naval, aerospace and mechanical engineers as well as other fields” (Prof. Patricio A. A. Laura, Prof. Carlos A. Rossit, Prof. Diana V. Bambill, Universidad Nacional del Sur, Argentina, Ocean Engineering). – “This book is an outstanding research monograph … extremely well written, informative, highly original … great scholarly contribution …. There is no comparable book discussing the combination of optimization and anti-optimization … magnificent monograph …. This book, which certainly is written with love and passion, is the first of its kind in applied mechanics literature, and has the potential of having a revolutionary impact on both uncertainty analysis and optimization” (Prof. Izuru Takewaki, Kyoto University, Engineering Structures). – “This book is a collection of a surprisingly large number of closed form solutions, by the author and by others, involving the buckling of columns and beams, and the vibration of rods, beams and circular plates. The structures are, in general, inhomogeneous. Many solutions are published here for the first time. The text starts with an instructive review of direct, semi-inverse, and inverse eigenvalue problems. Unusual closed form solutions of column buckling are presented first, followed by closed form solutions of the vibrations of rods. Unusual closed form solutions for vibrating beams follow. The influence of boundary conditions on eigenvalues is discussed. An entire chapter is devoted to boundary conditions involving guided ends. Effects of axial loads and of elastic foundations are presented in two separate chapters. The closed form solutions of circular plates concentrate on axisymmetric vibrations. The scholarly effort that produced this book is remarkable” (Prof. Werner Soedel, then Editor-in-Chief of Journal Sound and Vibration). – “The field has been brilliantly presented in book form …” (Prof. Luis A. Godoy et al., Institute of Advanced Studies in Engineering and Technology, Science Research Council of Argentina and National University of Cordoba, Argentina, Thin-Walled Structures).

xxii

Modern Trends in Structural and Solid Mechanics 1

– “Elishakoff is one of the pioneers in the use of the probabilistic approach for studying imperfection-sensitive structures” (Prof. Chiara Bisagni and Dr. Michela Alfano, Delft University of Technology; AIAA Journal). – “Recently, Elishakoff et al. presented an excellent literature review on the historical development of Timoshenko’s beam theory” (Prof. Zhenlei Chen et al., Journal of Building Engineering). Professor Isaac Elishakoff is the author or co-author of an impressive list of seminal books in the field of deterministic and non-deterministic mechanics, presented below. Books by Elishakoff Ben-Haim, Y. and Elishakoff, I. (1990). Convex Models of Uncertainty in Applied Mechanics. Elsevier, Amsterdam. Cederbaum, G., Elishakoff, I., Aboudi, J., Librescu, L. (n.d.). Random Vibration and Reliability of Composite Structures. Technomic, Lancaster. Elishakoff, I. (1983). Probabilistic Methods in the Theory of Structures. Wiley, New York. Elishakoff, I. (1999). Probabilistic Theory of Structures. Dover Publications, New York. Elishakoff, I. (2004). Safety Factors and Reliability: Friends or Foes? Kluwer Academic Publishers, Dordrecht. Elishakoff, I. (2005). Eigenvalues of Inhomogeneous Structures: Unusual Closed-Form Solutions of Semi-Inverse Problems. CRC Press, Boca Raton. Elishakoff, I. (2014). Resolution of the Twentieth Century Conundrum in Elastic Stability. World Scientific/Imperial College Press, Singapore. Elishakoff, I. (2017). Probabilistic Methods in the Theory of Structures: Random Strength of Materials, Random Vibration, and Buckling. World Scientific, Singapore. Elishakoff, I. (2018). Probabilistic Methods in the Theory of Structures: Solution Manual to Accompany Probabilistic Methods in the Theory of Structures: Problems with Complete, Worked Through Solutions. World Scientific, Singapore. Elishakoff, I. (2020). Dramatic Effect of Cross-Correlations in Random Vibrations of Discrete Systems, Beams, Plates, and Shells. Springer Nature, Switzerland. Elishakoff, I. (2020). Handbook on Timoshenko-Ehrenfest Beam and Uflyand-Mindlin Plate Theories. World Scientific, Singapore. Elishakoff, I. and Ohsaki, M. (2010). Optimization and Anti-Optimization of Structures under Uncertainty. Imperial College Press, London.

Preface

xxiii

Elishakoff, I. and Ren, Y. (2003). Finite Element Methods for Structures with Large Stochastic Variations. Oxford University Press, Oxford. Elishakoff, I., Lin, Y.K., Zhu, L.P. (1994). Probabilistic and Convex Modeling of Acoustically Excited Structures. Elsevier, Amsterdam. Elishakoff, I., Li, Y., Starnes Jr., J.H. (2001). Non-Classical Problems in the Theory of Elastic Stability. Cambridge University Press, Cambridge. Elishakoff, I., Pentaras, D., Dujat, K., Versaci, C., Muscolino, G., Storch, J., Bucas, S., Challamel, N., Natsuki, T., Zhang, Y., Ming Wang, C., Ghyselinck, G. (2012). Carbon Nanotubes and Nano Sensors: Vibrations, Buckling, and Ballistic Impact. ISTE Ltd, London, and John Wiley & Sons, New York. Elishakoff, I., Pentaras, D., Gentilini, C., Cristina, G. (2015). Mechanics of Functionally Graded Material Structures. World Scientific/Imperial College Press, Singapore.

Books edited or co-edited by Elishakoff Ariaratnam, S.T., Schuëller, G.I., Elishakoff, I. (1988). Stochastic Structural Dynamics – Progress in Theory and Applications. Elsevier, London. Casciati, F., Elishakoff, I., Roberts, J.B. (1990). Nonlinear Structural Systems under Random Conditions. Elsevier, Amsterdam. Chuh, M., Wolfe, H.F., Elishakoff, I. (1989). Vibration and Behavior of Composite Structures. ASME Press, New York. David, H. and Elishakoff, I. (1990). Impact and Buckling of Structures. ASME Press, New York. Elishakoff, I. (1999). Whys and Hows in Uncertainty Modeling. Springer, Vienna. Elishakoff, I. (2007). Mechanical Vibration: Where Do We Stand? Springer, Vienna. Elishakoff, I. and Horst, I. (1987). Refined Dynamical Theories of Beams, Plates and Shells and Their Applications. Springer Verlag, Berlin. Elishakoff, I. and Lin, Y.K. (1991). Stochastic Structural Dynamics 2 – New Applications. Springer, Berlin. Elishakoff, I. and Lyon, R.H. (1986), Random Vibration-Status and Recent Developments. Elsevier, Amsterdam. Elishakoff, I. and Seyranian, A.P. (2002). Modern Problems of Structural Stability. Springer, Vienna. Elishakoff, I. and Soize, C. (2012). Non-Deterministic Mechanics. Springer, Vienna. Elishakoff, I., Arbocz, J., Babcock Jr., C.D., Libai, A. (1988). Buckling of Structures: Theory and Experiment. Elsevier, Amsterdam.

xxiv

Modern Trends in Structural and Solid Mechanics 1

Lin, Y.K. and Elishakoff, I. (1991). Stochastic Structural Dynamics 1 – New Theoretical Developments. Springer, Berlin. Noor, A.K., Elishakoff, I., Hulbert, G. (1990). Symbolic Computations and Their Impact on Mechanics. ASME Press, New York.

Figure P.11. Elishakoff with his wife, Esther Elisha, M.D., during an ASME awards ceremony

On behalf of all the authors of this book, including those friends who were unable to contribute, we wish Prof. Isaac Elishakoff many more decades of fruitful works and collaborations for the benefit of world mechanics, in particular. Modern Trends in Structural and Solid Mechanics 1 – the first of three separate volumes that comprise this book – presents recent developments and research discoveries in structural and solid mechanics, with a focus on the statics and stability of solid and structural members. The book is centered around theoretical analysis and numerical phenomena and has broad scope, covering topics such as: buckling of discrete systems (elastic chains, lattices with short and long range interactions, and discrete arches), buckling of continuous structural elements including beams, arches and plates, static investigation of composite plates, exact solutions of plate problems, elastic and inelastic buckling, dynamic buckling under impulsive loading, buckling and post-buckling investigations, buckling of conservative and non-conservative systems, buckling of micro and macro-systems. The engineering applications

Preface

xxv

concern both small-scale phenomena with micro and nano-buckling up to large-scale structures, including the buckling of drillstring systems. Each of the three volumes is intended for graduate students and researchers in the field of theoretical and applied mechanics. Prof. Noël CHALLAMEL Lorient, France Prof. Julius KAPLUNOV Keele, UK Prof. Izuru TAKEWAKI Kyoto, Japan February 2021

1 Static Deformations of Fiber-Reinforced Composite Laminates by the Least-Squares Method

Accurate solutions of the linear elasticity equations governing three-dimensional static deformations of fiber-reinforced composite laminates are needed to efficiently design them for structural applications, by considering failure modes in them. In this chapter, the governing equations are written as first-order partial differential equations for three displacements, three transverse stresses and three in-plane strains. The mixed formulation facilitates the satisfaction of continuity conditions at an interface between two adjacent plies. These nine equations are numerically solved by minimizing residuals in them and in the boundary conditions by using the least-squares method. The functional of the residuals is evaluated by expressing the unknown quantities as the product of complete polynomials of different orders in the three independent coordinates and using appropriate quadrature rules. Minimization of the functional, with respect to coefficients appearing in the polynomials for the solution variables, provides a system of simultaneous linear algebraic equations that are numerically solved. It is shown that polynomial functions of degree, at most, 8 in the in-plane coordinates and 3 in the thickness coordinate for each layer of a laminate, provide accurate solutions for stresses and displacements when compared with the analytical solutions of problems. Through-the-thickness stress distributions are found to agree very well with those found analytically. It provides an efficient numerical scheme for accurately finding stresses and displacements in a laminate.

Chapter written by Devin BURNS and Romesh C. BATRA. Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

2

Modern Trends in Structural and Solid Mechanics 1

1.1. Introduction Composites are being increasingly used in transportation, aviation and defense industries because of their high specific strength and our ability to tailor their elastic moduli and ultimate strengths in desired directions and at critical locations in the structure. They are usually in the form of a laminate composed of numerous plies, with each ply having unidirectional fibers. The volume fraction of fibers in different plies can be varied to suit the intended application. Recognizing that design is an iterative process, it is imperative that we can easily and accurately analyze a structure’s response to external stimuli. Closed-form solutions for three-dimensional linear elasticity equations for composite laminates are scarce. Assuming that the Euler–Bernoulli beam theory applies to a composite laminated beam and the Kirchhoff–Love theory is relevant for a composite laminated plate, enables us to analytically solve some boundary value problems (see, for example, Jones (1998) and Hyer (2009)). Elishakoff has reviewed various theories for plates and laminates and has provided their historical perspective (Elishakoff 2018). However, for general structures, we resort to numerical methods for solving the governing equations. A challenge in solving problems for laminates is accurately finding transverse stresses that can cause delamination between adjacent plies. For simply supported edges, the linear elasticity equations have been solved by expressing the three displacements as double Fourier series in the in-plane coordinates and deducing ordinary differential equations for them in the thickness direction (Vlasov 1957; Pagano 1969; Srinivas and Rao 1970). For other boundary conditions at the edges, Vel and Batra employed the Eshelby–Stroh formalism and satisfied boundary conditions at the edges in the sense of Fourier series (Vel and Batra 1999; Vel and Batra 2000) that rely on St. Venant’s principle. This principle states that the solution at points away from the edges is unaffected by the boundary conditions, provided that they are equipollent to the same resultant force and moment (Toupin 1965). The distance from the edges, usually called the decay length, where the solution is unaffected, depends on the lowest frequency of free vibrations of a slice of the structure of characteristic length l and the maximum eigenvalue of the elasticity matrix. For composite laminates, the decay length has been estimated in Horgan and Baxter (1998). A common approach, called the state space method, is to take the three transverse stresses and the three displacements as variables, cast governing equations for them as first-order partial differential equations (PDEs) in the thickness coordinate, z, and find their values simultaneously. The boundary value problem is thus reduced to one that looks like an initial value problem. One way to solve these PDEs is to find a propagation matrix that relates variables for two

Static Deformations of Fiber-Reinforced Composite Laminates

3

different values of z (see, for example, Bahar (1975)). Kant and Ramesh have presented Goldberg and Bogdanoff’s method of numerically solving these six first-order equations (Goldberg and Bogdanoff 1957; Kant and Ramesh 1981) that satisfy the prescribed boundary conditions at the top and the bottom surfaces of a plate. Here, we also solve for the three in-plane strains, along with the aforementioned six unknowns, use the least-squares method, polynomial basis functions of possibly different orders in each one of the three directions defined on the entire structure’s domain, include in the functional to be minimized boundary conditions at the six edges of a laminate and derive a set of simultaneous algebraic equations; see, for example, Moleiro et al. (2011) who have analyzed numerous problems. The contribution of this work is in demonstrating that polynomial basis functions defined on the entire domain enable us to find a reasonably accurate solution of the linear elasticity equations, for rectangular composite laminates of side length equal to twice the thickness and sandwich structures with the ratio of the axial elastic modulus of a facesheet to that of the core less than 1000. The rest of the chapter is organized as follows. The problem formulation is described in section 1.2. In section 1.3, the developed algorithm is verified by comparing predictions from it with the analytical solutions for a simply supported four-layer laminate and a sandwich structure, which also has results for a three-layer laminate with two opposite edges simply supported and the other two edges either free or clamped. Significant outcomes of the work are briefly summarized as conclusions. 1.2. Formulation of the problem We formulate the problem for a rectangular a x b x h lamina that can be easily extended to a laminate made of p layers. We employ rectangular Cartesian coordinates (x1, x2, x3), shown in Figure 1.1, to describe the position of a material particle in the unstressed reference configuration. In the absence of body forces, infinitesimal deformations of the lamina are governed by the following equilibrium equations [1.1], stress–strain relations [1.2] and strain–displacement relations [1.3]: ,



=0

= =

[1.1]

, ,

+

,

=

=

, , , , = 1,2,3.

[1.2] [1.3]

4

Modern Trends in Structural and Solid Mechanics 1

Figure 1.1. Geometry and coordinate-axes of a laminated plate

In equations [1.1]–[1.3], σ is the stress tensor, e is the infinitesimal strain tensor, u is the displacement vector and C is the matrix of elasticities. Here, a layer is modeled as a transversely isotropic material with the axis of transverse isotropy along the fiber. A repeated index implies summation over the range of the index. The substitution from equation [1.3] into [1.2], and the result into equation [1.1] gives three second-order partial differential equations for the three displacement components that are to be solved under the prescribed boundary conditions of surface tractions on one part of the boundary and displacements on the other part. Of course, linearly independent components of u and the surface traction vector, fi = σij nj, can alternatively be prescribed at a point of the boundary. Here, nj is the jth component of the outward unit normal to the boundary. As is often done, we use the Voigt notation to express σ and e as six-dimensional vectors and write equation [1.2] as =

, ,

= 1,2,3,4,5,6.

[1.4]

In equation [1.4], C is a 6 x 6 symmetric matrix of elasticities of the layer material, and (σ1, σ2, σ3, σ4, σ5, σ6) = (σ11, σ22, σ12, σ13, σ23, σ33). A similar notation is used for e. We use a mixed formulation and take sk = (u1k, u2k, u3k, σ4k, σ5k, σ6k, e1k, e2k, e3k) as unknowns at a point in each plate layer. In order to solve for sk, we define the following residuals on the kth layer that only involve first-order derivatives for the elements of sk: ≡

,





,



[1.5a]

Static Deformations of Fiber-Reinforced Composite Laminates



,

+

,

−2





,



,



,



,





,



,



,



,





=



+

+ =

+

, ,

+

,

+

,

,

,

,

[1.5c] ,

+

+

,

+

,

+

,

+

[1.5b]

,

+

,

+

,



+

,

+

,

+

,

+

,

=



+

,

5

+

,

,

,

In equations [1.5b] and [1.5c], the coefficients represent the elasticities of the kth layer material, with respect to the layer material principal axes rotated with respect to the analysis coordinate system (see Figure 1.1). The quantities are found by expressing the in-plane stresses and transverse strains in terms of the transverse stresses and in-plane strains (i.e. the layerwise continuous variables) from equations [1.2] and [1.3] (see Moleiro et al. 2011). It should be highlighted that in-plane strains have been incorporated into the mixed formulation, in order to recast the governing equations into first-order form, which allows the use of C0 basis functions for the interpolation of unknown variables, discussed later. While this increases the number of unknowns to solve for, the advantage is that continuity of surface tractions and displacements at layer interfaces can now be enforced. = 0, a = 1, 2, …, 9 are the nine equations for the nine We note that unknowns in sk. Recall that there are three boundary conditions prescribed at each bounding face of the lamina. Generally, the top and the bottom faces of a lamina have prescribed tractions, = , on them, and edges x = 0, a and y = 0, b have a combination of and , where and are the known function of the in-plane coordinates. For example, at a clamped edge, = 0, and at a free edge, = 0. At a simply supported edge x = 0, we set = 0,

= 0,

= 0.

[1.6]

6

Modern Trends in Structural and Solid Mechanics 1

Equation [1.6]3 is equivalent to a null in-plane normal stress at the edge x = 0. Since in-plane stresses are not directly computed in the present formulation, we write this boundary condition residual in terms of the variables in s as =

+

+

+

[1.7]

which also uses the constitutive relation in equation [1.4]. Thus, for each one of the six bounding faces, we will have three residuals that we denote by R10 through R27. We define the functional, , in terms of the residuals, with contributions from each material layer as = ( ∑

Ω+∑





Ω)

[1.8]

We note that the repeated index a goes from 1 to 9, and the second surface (b = 10, 11, …, 27) integral is for the six bounding surfaces of a lamina, and in it, are residuals for the boundary conditions of the kth layer. In equation [1.8], Ω denotes either the in-plane domain of the plate or one of its edge surfaces, hk is the thickness of the kth layer. In evaluating integrals with respect to x3, elasticities of the individual layer are considered. Each element of the nine-dimensional vector sk is expressed as the product of complete Lagrange polynomials of degrees N1, N2 and N3 in x1, x2 and x3 defined as follows: =∑ ( )=





( )

(

)(

)

(

)

( )(

( ) )

( )

( )

[1.9] [1.10]

Note that equation [1.9] has 9 x N1 x N2 x N3 unknowns, , for each layer. In equation [1.10], the basis function is written in natural coordinates, , and in terms of the Pth-order Lagrange polynomial LP( ) and its derivative, indicated by the is a root of the equation Pn( ) = 0, where Pn is a prime symbol. The quantity Legendre polynomial of order n. Basis functions given in [1.10] are associated with Gauss–Lobatto points. Substitution from equation [1.10] into equation [1.9], the result into equation [1.8], and the numerical evaluation of the integral by using the in the three directions, gives as a Gauss–Lobatto quadrature rule of order . We deduce the needed linear algebraic equations by setting function of = 0.

[1.11]

Static Deformations of Fiber-Reinforced Composite Laminates

7

We realize that expressions for the residuals have different units. When equation [1.11] is written as KA = F, it is likely that the use of non-dimensional variables throughout the chapter will improve the condition number of the matrix K and reduce error. However, we have not tried this. A feature of the equations derived from [1.11] using basis functions of type [1.10] is that they are insensitive to shear-locking effects, which means that reduced integration is not needed in the thickness direction. We note that the above formulation holds for a laminate, when the continuity of variables u1, u2, u3, σ4, σ5, σ6, e1, e2, e3 is enforced by adding the appropriate residuals in equation [1.8] or using a layerwise theory. Here, we use a layerwise theory, where the contribution from each layer is included in the summation in equation [1.8] and the continuity of the variables in s at each layer interface is ensured. 1.3. Results and discussion 1.3.1. Verification of the numerical algorithm To verify the algorithm and to establish the accuracy of computed results, we study the problem analytically analyzed by Pagano (1969). It involves a four-layered [0/90/90/0] simply supported square laminate of side length a, with the sinusoidal surface traction = [0,0, −

sin

sin

]

[1.12]

applied only on the top surface. The material of the layers has the following values of the moduli: = 25,

= 0.5,

= 0.2,

=

= 0.25

[1.13]

Here, E, G and denote Young’s modulus, shear modulus and Poisson’s ratio, respectively, and subscripts L and T indicate directions parallel and transverse to the fiber direction. Following Pagano, we express the results in terms of the non-dimensionalized quantities defined in equation [1.4] and employ (x, y, z) = (x1, x2, x3) as the coordinate axes and (u, v, w) = (u1, u2, u3) , ,

,

= =

( ⁄ )

1 ( ( ⁄ℎ)

,

,

,

,

) =

[1.14]

8

Modern Trends in Structural and Solid Mechanics 1

( , ̅) =

ℎ( ⁄ℎ)

( , ),

=

100 ℎ( ⁄ℎ)

.

In Table 1.1, we compare our results of select quantities with those in Pagano (1969) for the plate aspect ratio a/h = 100, 10, 4 and 2. It is clear that the developed least-squares method algorithm yields highly accurate results for the simply supported laminate. a/h

-

⎯σxx 1

⎯σyy 2

⎯σxy 3

⎯σxz 4

⎯σyz 5

⎯w 6

Pagano

1.38841 -0.91165

0.83508 -0.79465

-0.08630 0.06732

0.15300

0.29458

5.0745

Present

1.38020 -0.90607

0.83038 -0.79049

-0.08599 0.06711

0.15311

0.29428

5.0643

Pagano

0.72026 -0.68434

0.66255 -0.66551

-0.04666 0.04581

0.21933

0.29152

1.93672

Present

0.72020 -0.68427

0.66246 -0.66541

-0.04665 0.04575

0.21939

0.29154

1.93660

Pagano

0.55861 -0.55909

0.40095 -0.40257

-0.02750 0.02764

0.30137

0.19595

0.73698

Present

0.55862 -0.55910

0.40096 -0.40257

-0.02747 0.02761

0.30140

0.19597

0.73698

Pagano

0.53885 -0.53887

0.27101 -0.27103

-0.02135 0.02136

0.33880

0.13894

0.43460

Present

0.53883 -0.53885

0.27100 -0.27102

0.021353 -0.021355

0.33880

0.13894

0.43460

2

4

10

100

Table 1.1. Comparison of the results with the 3D exact solution of Pagano for the [0/90/90/0] laminate

1 – (a/2, a/2, ±h/2)

2 – (a/2, a/2, ±h/4)

5 – (a/2, 0, 0)

6 – (a/2, a/2, 0)

3 – (0, 0, ±h/2)

4 – (0, a/2, 0)

For a/h = 100, the maximum error in the computed quantities equals 0.023% for the in-plane shear stress at point (0,0, −ℎ/2), and for a/h = 2, the maximum error is 0.612% for the in-plane axial stress at point ( /2, /2, −ℎ/2). The errors for a/h = 10 and 4 are between those for a/h = 100 and 2. The through-the-thickness ( /2, /2, ), (0, /2, ), ( /2,0, ) and (0, /2, ), plots of ( /2, /2, ) coincide well with those from Pagano’s solution and are omitted.

Static Deformations of Fiber-Reinforced Composite Laminates

9

For a/h = 10 and 2, we conducted the following five numerical experiments, E1, E2, …, E5, by varying N1, N2 and N3. In E1 and E2, N3 was set equal to either 3 or 4 and N1 = N2 was assigned values 4, 6 and 8; in E3 and E4, N1 = N2 = 4 and N3 was given values 3, 4, 6 and 8; and in E5, N1 = N2 = N3 were assigned values 4, 6 and 8. Thus, a total of 13 combinations of N1, N2 and N3 were evaluated for each value of , , and at typical points were compared a/h, and values of u, v, w, with their analytical values. For a/h = 10, N1 = N2 = 8, N3 = 3 (total degrees of freedom (DOFs) = 9,477) provided an error of less than 0.1% in the value of each of these seven variables. Keeping N1 = N2 = 8 and increasing N3 = 4, 5 marginally decreased the error, suggesting that there is no benefit derived from the increased computation cost. For a/h = 4, all 13 combinations of values of N1, N2 and N3 decreased from 37.7% for resulted in an error of 7.25% in w and the error in N1=N2=N3 = 4 to 0.03% for N1 = N2 = N3 = 8 (DOFs = 24,057). 1.3.2. Simply supported sandwich plate We now study deformations of a sandwich plate with the top and bottom facesheets made of an orthotropic Aragonite crystal and the core made of a softer material. Material properties of the facesheets and the core are the same as those in Srinivas and Rao (1970). The top surface of the plate is loaded by sinusoidal tractions given by equation [1.12] and the bottom surface is traction-free. Since the load on the top surface given by equation [1.12] is different from that in Srinivas and Rao (1970), we first solved the problem analytically using their methodology, then verified our software by ensuring that values of displacements and stresses at several points agreed with those reported in Srinivas and Rao (1970). The thickness of each facesheet equals 0.1h and that of the core 0.8h, where h equals the sandwich plate thickness. Defining β = Ex1/Ex2, where Ex1 and Ex2 are, respectively, elastic moduli of the facesheets and the core in the x direction, we compare the results from the least-squares method and the analytical solution for β = 1, 10, 100, 1000, 10,000, and aspect ratio a/h = 10, 5 and 2 in Tables 1.2–1.4. Quantities compared are the , and , normalized as wEx2/q0, σxx/q0, σyy/q0 centroidal deflection, stresses and σxz/q0. Note that z = 0.4h+ and 0.4h-, respectively, represent interfaces between the top facesheet and the core. For the three aspect ratios and for all values of β except β = 10,000, the least-squares methodology gives results that differ from their analytical values by less than 0.4%. Even for β = 10,000, the maximum error in all quantities is less than 1.3%, except for that in σxz. Thus, we should not use the least-squares method for β = 10,000. We have not investigated if increasing values of N1, N2 and N3 in equation [1.9] will reduce this error.

10

Modern Trends in Structural and Solid Mechanics 1

For a/h = 2 and β = 1000, the error in the 15 variables listed in Table 1.4 was less than 1% for N1 = N2 = 6 and N3 = 4 (DOFs = 7,497) and less than 0.6% for N1 = N2 = 8 and N3 = 4 (DOFs = 12,393). Thus, the present method provides accurate values of the variables at critical points for a rather modest computational cost. 1.3.3. Laminate with arbitrary boundary conditions We now present results for a [0/90/0] laminate simply supported on edges y = 0, b, with the other two edges either clamped or traction-free, and loaded on the top surface by the surface tractions listed in equation [1.12]. For the plate with two edges free, the present results have been computed with N1 = N2 = 11 and N3 = 5 in equation [1.9] for a total of 14,157 DOFs. These values could not be increased due to memory limitation of the Dell laptop used for the computational work. We are now converting the code from MATLAB to C++, which will enable us to use larger values of N1, N2 and N3. In Table 1.5, the numerical results for different quantities for the laminates with a/h = 5 and 10 are compared with their values reported by Vel and Batra (1999), who used the Eshelby–Stroh formalism. The maximum difference of 4.2% in the two sets of values of the six quantities suggests that the least-squares method provides a reasonably accurate solution for this problem.

⎯W ⎯σxx

β→

1

10

100

1000

10,000

(0.5, 0.5, -0.5h)

-436.824

-102.422

-32.9858

-21.6946

-10.5531

(-436.820)

(-102.422)

(-32.9858)

(-21.6942)

(-10.5452)

-24.6675

-45.1109

-52.1977

-155.576

-694.701

(-24.6687)

(-45.1110)

(-52.1980)

(-155.574)

(-694.429)

-19.4585

-32.7715

-11.4111

116.821

676.521

(-19.4537)

(-32.7713)

(-11.4113)

(116.819)

(676.846)

-19.4585

-3.29448

-0.13298

0.09864

0.05351

(-19.4537)

(-3.29446)

(-0.13300)

(0.09864)

(0.05349)

19.3977

3.26259

0.11506

-0.11458

-0.06898

(19.3929)

(3.26259)

(0.11507)

(-0.11457)

(-0.06911)

19.3977

32.6356

11.6322

-112.483

-631.675

(19.3929)

(32.6355)

(11.6324)

(-112.471)

(-632.023)

24.5962

44.8809

51.4978

150.743

649.367

(24.5973)

(44.8810)

(51.4980)

(150.733)

(650.206)

(0.5, 0.5, z) z = 0.5h

z = 0.4h+

z = 0.4h-

z = -0.4h+

z = -0.4h-

z = -0.5h

Static Deformations of Fiber-Reinforced Composite Laminates

β→ ⎯σyy

1

10

100

1000

10,000

-15.6259

-30.7654

-48.9480

-123.891

-442.713

(-15.6264)

(-30.7654)

(-48.9478)

(-123.889)

(-442.514)

-12.4002

-23.1401

-23.8576

43.4841

399.777

(-12.3978)

(-23.1399)

(-23.8573)

(43.4832)

(399.971)

-12.4002

-2.47206

-0.41064

-0.12230

-0.08903

(-12.3978)

(-2.47205)

(-0.41064)

(-0.12230)

(-0.08946)

12.3427

2.33393

0.23404

-0.05935

-0.09017

(12.3404)

(2.33392)

(0.23404)

(-0.05934)

(-0.09129)

12.3427

23.4274

24.5512

-40.2356

-371.648

(12.3404)

(23.4273)

(24.5509)

(-40.2337)

(-374.066)

15.5620

30.9947

49.0753

121.504

415.435

(15.5625)

(30.9947)

(49.0751)

(121.502)

(417.599)

0.99214

1.78299

1.64940

1.24653

0.61880

(0.99210)

(1.78301)

(1.64946)

(1.24664)

(0.59365)

0.99152

1.78239

1.64878

1.24591

0.61820

(0.99151)

(1.78241)

(1.64885)

(1.24609)

(0.72700)

(0.5, 0.5, z) z = 0.5h

z = 0.4h+

z = 0.4h-

z = -0.4h+

z = -0.4h-

z = -0.5h

⎯σxz

11

(x=0, y=0.5) z = 0.4h+

z = -0.4h+

Table 1.2. Normalized results for the sandwich plate of the aspect ratio a/h = 10. Values in the upper (lower) row are from the analytical (least-squares) solution

⎯W ⎯σxx

β→

1

10

100

1000

10,000

(0.5, 0.5, -0.5h)

-32.8669

-11.0284

-5.97592

-3.72388

-0.75513

(-32.8636)

(-11.0283)

(-5.97592)

(-3.72393)

(-0.75586)

-6.24479

-11.5195

-22.3767

-111.720

-325.689

(-6.24849)

(-11.5194)

(-22.3764)

(-111.718)

(-325.756)

-4.71502

-5.97928

10.0974

104.342

323.587

(-4.69630)

(-5.97819)

(10.0971)

(104.339)

(323.515)

(0.5, 0.5, z) z = 0.5h

z = 0.4h

+

12

Modern Trends in Structural and Solid Mechanics 1

β→

1

10

100

1000

10,000

z = 0.4h-

-4.71502

-0.61521

0.08244

0.08837

0.02295

(-4.69630)

(-0.61510)

(0.08244)

(0.08836)

(0.02297)

4.63045

0.58473

-0.08814

-0.09061

-0.02446

(4.61255)

(0.58471)

(-0.08813)

(-0.09061)

(-0.02461)

4.63045

5.85733

-8.65984

-86.8644

-182.172

(4.61255)

(5.85718)

(-8.65970)

(-86.8631)

(-183.119)

6.11899

11.0239

20.1365

93.4349

183.620

(6.12250)

(11.0236)

(20.1366)

(93.4337)

(184.363)

-4.28086

-9.36617

-20.0972

-74.2391

-201.621

(-4.28249)

(-9.36611)

(-20.0970)

(-74.2373)

(-201.636)

-3.31557

-5.92194

-0.10678

58.6449

197.676

(-3.30643)

(-5.92142)

(-0.10682)

(58.6434)

(197.642)

-3.31557

-0.74984

-0.17007

-0.08703

-0.06604

(-3.30643)

(-0.74979)

(-0.17007)

(-0.08704)

(-0.06579)

3.24276

0.61151

0.00125

-0.08155

-0.06794

(3.23401)

(0.61150)

(0.00124)

(-0.08154)

(-0.06842)

3.24276

6.20664

1.52702

-47.3508

-110.275

(3.23401)

(6.20655)

(1.52704)

(-47.3527)

(-110.197)

4.18246

9.41935

19.2533

63.5372

114.682

(4.18400)

(9.41922)

(19.2532)

(63.5391)

(114.477)

0.49637

0.83806

0.71375

0.47356

0.12573

(0.49615)

(0.83804)

(0.71377)

(0.47363)

(0.10374)

0.49166

0.82533

0.69746

0.45722

0.11274

(0.49144)

(0.82532)

(0.69747)

(0.45722)

(0.12017)

+

z = -0.4h

z = -0.4h-

z = -0.5h

⎯σyy

(0.5, 0.5, z) z = 0.5h

z = 0.4h

+

z = 0.4h+

z = -0.4h

z = -0.4h-

z = -0.5h

⎯σxz

(X=0, Y=0.5) z = 0.4h+ +

z = -0.4h

Table 1.3. Normalized results for the sandwich plate of the aspect ratio a/h = 5. Values in the upper (lower) row are from the analytical (least-squares) solution

Static Deformations of Fiber-Reinforced Composite Laminates

13

β→

1

10

100

1000

10,000

⎯W

(0.5, 0.5, -0.5h)

-1.55045

-0.80673

-0.47167

-0.07767

-0.00174

(-1.55041)

(-0.80672)

(-0.47167)

(-0.07767)

(-0.00179)

⎯σxx

(0.5, 0.5, z) -1.30381 (-1.30409) -0.71151 (-0.70956) -0.71151 (-0.70956) 0.53608 (0.53463) 0.53608 (0.53463) 0.90503 (0.90518)

-3.68270 (-3.68274) 0.70904 (0.70917) 0.05412 (0.05413) -0.03570 (-0.03569) -0.34646 (-0.34641) 1.92738 (1.92735)

-17.8693 (-17.8692) 15.5178 (15.5175) 0.13952 (0.13952) -0.06779 (-0.06780) -6.56097 (-6.56090) 7.39749 (7.39739)

-63.5162 (-63.5153) 62.5785 (62.5774) 0.05564 (0.05564) -0.01411 (-0.01411) -11.5108 (-11.5113) 11.6037 (11.6042)

-92.9483 (-92.9222) 92.7600 (92.7827) 0.00816 (0.00814) -0.00082 (-0.00045) -2.59311 (-1.9654) 2.58611 (2.41811)

-1.12800 (-1.12813) -0.73700 (-0.73607) -0.73700 (-0.73607) 0.58971 (0.58899) 0.58971 (0.58899) 0.83614 (0.83621)

-3.06623 (-3.06624) -0.31380 (-0.31374) -0.18445 (-0.18444) 0.06560 (0.06560) 0.75187 (0.75190) 2.17320 (2.17319)

-11.9076 (-11.9075) 8.78405 (8.78389) -0.05495 (-0.05496) -0.04952 (-0.04952) -2.96035 (-2.96026) 5.68229 (5.68223)

-39.5773 (-39.5766) 38.4871 (38.4863) -0.02475 (-0.02477) -0.03041 (-0.03040) -6.72640 (-6.72638) 7.57969 (7.57976)

-57.5749 (-57.5841) 57.3827 (57.3665) -0.00442 (-0.00463) -0.00524 (-0.00247) -1.55007 (-1.46223) 1.65537 (1.42636)

0.22882 (0.22878) 0.17656 (0.17652)

0.36501 (0.36502) 0.23854 (0.23853)

0.30164 (0.30166) 0.16531 (0.16530)

0.11080 (0.11086) 0.03726 (0.03723)

0.01593 (0.03117) 0.00302 (0.00056)

z = 0.5h z = 0.4h+ z = 0.4hz = -0.4h+ z = -0.4hz = -0.5h ⎯σyy

(0.5, 0.5, z) z = 0.5h z = 0.4h+ z = 0.4hz = -0.4h+ z = -0.4hz = -0.5h

⎯σxz

(X=0, Y=0.5) z = 0.4h+ z = -0.4h+

Table 1.4. Normalized results for the sandwich plate of the aspect ratio a/h = 2. Values in the upper (lower) row are from the analytical (least-squares) solution

14

Modern Trends in Structural and Solid Mechanics 1

a/h

5

10

Variable

Clamped–Clamped

Free–Free

Vel and Batra

Present

Vel and Batra

Present

⎯W (a/2, b/2, h/2)

1.1800

1.1771

1.5250

1.4649

⎯σxx (a/2, b/2, 0)

-4.2350

-4.2757

-6.9870

-7.2402

⎯σxx (a/2, b/2, h)

4.5040

4.5437

7.1800

7.4300

⎯σyy (a/2, b/2, h/3)

-3.7260

-3.7185

-4.784

-4.582

⎯σyy (a/2, b/2, 2h/3)

3.5760

3.5652

4.639

4.4364

⎯σyz (a/2, 0, h/2)

1.4700

1.4711

1.9110

1.8494

⎯W (a/2, b/2, h/2)

0.4460

0.4457

0.7530

0.72872

⎯σxx (a/2, b/2, 0)

-3.0000

-2.9746

-5.8980

-5.983

⎯σxx (a/2, b/2, h)

3.0320

3.0066

5.9060

5.9901

⎯σyy (a/2, b/2, h/3)

-1.7130

-1.7113

-2.882

-2.7854

⎯σyy (a/2, b/2, 2h/3)

1.6740

1.672

2.845

2.748

⎯σyz (a/2, 0, h/2)

0.7220

0.7243

1.2280

1.1973

Table 1.5. Normalized results for the [0/90/0] laminated plate with C–C–SS–SS and F–F–SS–SS boundary conditions

1.4. Remarks This work differs from the classical Ritz method in that the basis functions are not required to satisfy the essential boundary conditions. In the Ritz method, we can use a penalty parameter to enforce essential boundary conditions in a weak sense and use the present basis functions. In the traditional finite element method (FEM), we discretize each layer into disjoint domains called finite elements (FEs) and the compact support of the FE basis function for a node equals all FEs sharing that node. We could use the FEM by first taking the inner product of the nine governing equations Ra = 0 with a nine-dimensional test function and integrating the result over the laminate. The matrix K in K A = F will not be symmetric. We can improve on the accuracy of the numerical solution by either increasing the order of polynomials in the basis functions or reducing the element size or both. Even though we have not tried it, in general, it takes more computational resources than those needed for the present least-squares approach. Of course, only using one FE with the current basis functions is possible. Then, the difference will be in deriving matrices K and F and satisfying boundary conditions.

Static Deformations of Fiber-Reinforced Composite Laminates

15

1.5. Conclusion We have numerically solved three-dimensional linear elasticity equations by taking the three transverse stresses, three strain–displacement relations in the xy plane and three displacement components as independent variables and using the least-squares method to minimize the residuals in the expressions for these nine variables. For a simply supported rectangular plate, using complete polynomials of degree 4 in the z direction, and at most, 10 in each of the x and y directions provides, with very few degrees of freedom, an accurate solution relative to the exact solutions with error less than 0.3% in each of the nine variables, even for a very thick plate of aspect ratio 2. However, when edges are either clamped or free, a higher degree polynomial is needed in each direction to achieve the same accuracy. For a sandwich plate with ratio β of the axial modulus of the facesheet to that of the core up to 1000 and aspect ratios varying from 2 to 100, accurate results are obtained for each one of the nine variables. For β = 10,000, the accuracy considerably deteriorates with errors approaching 100%. All of these computations have been done using a Dell laptop. Limitations of the work include satisfying governing equations and boundary conditions in the least-squares sense rather than pointwise. 1.6. Acknowledgments The authors gratefully acknowledge the partial support of this work by the Office of Naval Research through grants N00014-18-1-2548 and N00014-20-1-2876 to Virginia Polytechnic Institute and State University, with Dr. Y. D. S. Rajapakse as the Program Manager. The views expressed in this chapter are those of the authors and neither of the funding agency nor of their Institution. 1.7. References Bahar, L.Y. (1975). A state space approach to elasticity. Journal of the Franklin Institute, 299, 33–41. Elishakoff, I. (2019). Handbook on Timoshenko–Ehrenfest Beam and Uflyand–Mindlin Plate Theories. World Scientific, Singapore. Goldberg, J.E. and Bogdanoff, J.L. (1957). Midwest applied science corporation (MASC). Report, No. 57–7. Horgan, C.O. and Baxter, S.C. (1998). Saint-Venant’s principle for sandwich structures. Mechanics of Sandwich Structures, 113–122.

16

Modern Trends in Structural and Solid Mechanics 1

Hyer, M.W. (2009). Stress Analysis of Fiber-Reinforced Composite Materials. DEStech Publications, Inc., Lancaster, PA. Jones, R.M. (1998). Mechanics of Composite Materials. CRC Press, Boca Raton. Kant, T. and Ramesh, C.K. (1981). Numerical integration of linear boundary value problems in solid mechanics by segmentation method. International Journal for Numerical Methods in Engineering, 17, 1233–1256. Moleiro, F., Mota Soares, C.M., Mota Soares, C.A., Reddy, J.N. (2011). A layerwise mixed least-squares finite element model for static analysis of multilayered composite plates. Computers & Structures, 89, 1730–1742. Pagano, N.J. (1969). Exact solutions for composite laminates in cylindrical bending. Journal of Composite Materials, 3, 398–411. Srinivas, S. and Rao, A.K. (1970). Bending, vibration and buckling of simply supported thick orthotropic rectangular plates and laminates. International Journal of Solids Structures, 6, 1463–1481. Toupin, R.A. (1965). Saint-Venant’s principle. Archive for Rational Mechanics and Analysis, 18, 83–96. Vel, S.S. and Batra, R.C. (1999). Analytical solutions for rectangular thick laminated plates subjected to arbitrary boundary conditions. AIAA Journal, 37, 1464–1473. Vel, S.S. and Batra, R.C. (2000). The generalized plane strain deformations of thick anisotropic composite laminated plates. International Journal of Solids Structures, 37, 715–733. Vlasov, B.F. (1957). On one case of bending of rectangular thick plates. In Vestnik Moskovskogo Universiteta. Serieila Matematiki, Mekhaniki, Astronomii, Fiziki, Khimii, 2(2), 25–34.

2 Stability of Laterally Compressed Elastic Chains

2.1. Introduction Nominally identical elastic components are frequently arranged periodically in many natural and artificial structures. Examples include multilayer laminated composites (Dodwell and Hunt 2014), lattice structures with regularly placed struts (Cuan-Urquizo and Bhaskar 2018), structures with orderly placed holes and cut-outs in cold-formed rods, beams and plates (Moen and Schafer 2009a) and geological strata of rock and sedimentary soil (Hunt et al. 2006; Dodwell et al. 2012). While such structures often operate under mechanical compression and may be prone to buckling instability, most studies so far are confined to the apparent stiffness of periodic structures or wave propagation in such systems. This makes the prediction of critical buckling loads and associated mode shape of such periodic structures an important engineering task, which is what we will undertake here. Perfect spatial periodicity may not be present in many practical structures (natural or artificial). This necessitates the inclusion of spatial disorder within the analysis of structural instability. A disorder is any geometrical or physical irregularity that appears in the otherwise identical structural members. Such irregularities may include differences in geometric dimensions of the structural members, or in the material properties, such as elastic modulus, shear modulus, Poisson ratio, etc. Elastically coupled simple structural chains that are periodic in the direction of compression, such as axially compressed multi-span beams, have been studied before (Li et al. 1995; Xie 1995). The periodicity there is along the direction of loading. By contrast, here we consider a mechanical problem where nominally identical elastic structures

Chapter written by Andrii I AKOVLIEV, Srinandan DASMAHAPATRA and Atul B HASKAR.

18

Modern Trends in Structural and Solid Mechanics 1

may be coupled across their length, whereas loading is still applied longitudinally. A simple experiment may consist of stacks of sheets that are held together laterally while being compressed longitudinally. While such a simple experiment does not necessarily bring out the behavior of such elastically coupled structures quantitatively, it certainly is illustrative of the qualitative behavior. The essential feature of such elastically connected structures is periodicity across the bays and not along them. Here, we consider a simple chain model as a toy problem, where identical springs couple rigid rods characterized by small random irregularities, represented by small differences in length between the rods. We show that randomness localizes the buckling amplitude, which decays exponentially with distance from the impurity. We associate the localization with the ratio of the strength of the elastic coupling between rods to the magnitude of the variations in rod lengths or a disorder. We also show that the mathematical structure of the localization in the buckling amplitude in the chain and the multilayer structure is similar to wave localization, which was pioneered in the transport properties of conduction electrons in the presence of random impurities. There exists an extensive literature on wave propagation in periodic elastic media. The solutions to this class of problems are well known and are given by the so-called Floquet–Bloch theorem (Floquet 1883; Bloch 1929). The solutions represent physically extended propagation modes. If disorders exist in periodic systems, modes begin to localize spatially. The spatial localization of the modes is characterized by exponential decay – the rate of decay depends on the ratio of coupling to disorder. While this story is fairly well established in the field of wave propagation, an analogous phenomenon in the context of the stability of periodic elastic structures is less familiar. Moreover, within the few works concerning periodic elastic structures buckling, nearly all previous studies are about periodicity along the loading direction, such as axially compressed multi-span structures (Pierre and Plaut 1989; Li et al. 1995; Xie 1995; Xie and Ibrahim 2000). The simple chain considered here as an expository device has the features of being repetitive across bays, while the structure is compressed along the bays. A substantial body of literature exists on wave propagation in periodic media, such as electromagnetic waves in crystals (Brillouin 2003), waves in periodic elastic structures (Phani et al. 2006), the vibration of multi-span beams and chains of oscillators (Hodges and Woodhouse 1983; Castanier and Pierre 1995). In contrast to this, the static instability of structural assemblies of periodically coupled members has been, relatively, less studied. This chapter makes a contribution to this area and reports research into buckling behavior that parallels phenomena frequently encountered in vibration studies (Pierre and Plaut 1989; Elishakoff et al. 1995, 1997; Li et al. 1995; Xie 1995; Xie and Elishakoff 2000). Analysis of the available literature (Chai et al. 1981; Cho and Kim 2001; Rodman et al. 2008; Juh´asz and Szekr´enyes 2017) also suggests that studies focused on the buckling of layered structures such as laminated composites rarely make a connection with the studies of

Stability of Laterally Compressed Elastic Chains

19

buckling of periodic structures carried out by researchers from the wave propagation field. It turns out that there are close mathematical and physical analogies between the instability problem of coupled repetitive structures and those of the propagation of waves in periodic media. Here, we expand on these similarities by drawing upon the loss of wave propagation in lattices induced by random impurities (Anderson 1958) to investigate analogous phenomena in static instability problems. Indeed, we have identified only a few studies of the instability behavior of periodic structures. Pierre and Plaut (1989) studied the effect of irregularity on the instability of a two-span beam, while Xie (1995) studied multi-span disordered continuous beams using Lyapunov exponents as a measure of the localization of the buckling mode. Similarly, Li et al. (1995) studied buckling of axially loaded multi-span beams with the disorder in a single span. The authors find that the buckling amplitude exhibits weak or strong localization in the vicinity of a disordered segment depending on the strength of coupling between the spans, the magnitude of the disorder and the location of the disordered span. Elishakoff et al. (1995), Xie (1997), Li et al. (2005) and others studied the instability behavior of nominally periodic plates in the presence of disorder using a model of a long plate with the regularly positioned stiffeners along its length. The plate is typically compressed by loading P distributed along the shorter edges that are also simply supported, while the longer edges were considered free. Elishakoff et al. (1995) showed that, as the number of stiffeners increases, i.e. as the plate approaches perfect periodicity, it becomes more sensitive to disorder. Furthermore, as stiffeners naturally provide coupling between the neighboring spans of the plate, the critical buckling load and mode shape were shown to be sensitive to a single universal factor, namely, the ratio of the coupling strength to the degree of disorder. A similar plate but with many stiffeners was studied by Xie (1998) where the calculations were simplified by the introduction of a transfer matrix method. Transfer matrices simplify calculations for a large number of periodic structural arrays, and suitable approximations greatly reduce the number of degrees of freedom needed for computation. If the structure contains a large number of spans that are randomly disordered according to a known distribution, we may use Furstenberg’s theorem (Furstenberg 1963) in conjunction with the transfer matrix method to estimate the localization factors. Xie (1998) used the numerical approach based on the transfer matrix method to calculate the localization factors for the long but finite plate with N = 5 × 105 stiffeners. Xie and Ibrahim (2000) also employed the transfer matrix method to rib-stiffened plates. However, they formulate the transfer matrices using a finite strip method. Xie and Elishakoff (2000) also studied a similar problem of the stiffened plate using the Kantorovich approach. Nearly all these studies are limited to multi-span structures compressed along their length, the same as the direction of periodicity. By contrast, here we are motivated by the problem of axially loaded layered structures. We consider them as structures with periodicity across bays, not along them. A close analogy of the class

20

Modern Trends in Structural and Solid Mechanics 1

of problem considered here is that of vibration problems of bladed disks coupled by a hub – these blades are like bays, or layers, that can be considered as forming a chain, even though the physical context of such a problem is structural dynamics rather than elastic instability. Moreover, the problems studied in the current work involve relatively simple analytical models. This is to complement and provide insight into studies of industrially relevant structures that involve either extensive computational resources, such as the finite element method (Rikards et al. 2001; Moen and Schafer 2009b; Nguyen-Thoi et al. 2013), as well as mesh-free and semi-analytical methods (Byklum et al. 2004; Peng et al. 2006), or a direct experimental characterization (Cheng et al. 2013). While the experimental and large-scale numerical studies provide a valuable estimation of the stability of these periodic structures, it is difficult to extract meaningful interpretations on the role structural parameters play. Our study is inspired by the need to gain a simple analytic understanding of the generic instability behavior of this class of problems. In the next section, we describe an elementary experiment of layers of elastic sheets that are compressed while the small lateral coupling is provided by grips. 2.2. Compression of stacked elastic sheets We examine a structure consisting of several layers of nearly identical elastic sheets coupled at the ends experimentally. These sheets are stacked on top of each other and compressed while being held together. A schematic diagram of the structure and the applied loading is shown in Figure 2.1. We believe that the insights gained from this model will inform the stability of layered composites, strata and similar structures. The samples were made of cardboard sheets, which are much stiffer than paper. A single sheet of the size considered does not bend under its own weight. Such sheets are industrially produced, widely available and have excellent dimensional consistency. It is also easy to cut them into the desired shape, and they require only a few Newtons of loading to trigger buckling. Therefore, complex procedures of milling and then testing using custom-made grips on bulky experimental equipment designed for testing large and stiff samples are not needed. The coupling between the sheets can be simply achieved by gluing their ends to each other. What we seek to achieve with such structures is not their complete experimental characterization but to gain a qualitative idea of the buckling modes involved.

P

P

Figure 2.1. A schematic diagram of a structure consisting of identical layers stacked on top of each other and coupled at the ends. If the number of layers is large, the structure may be considered to approach periodicity in the stacking direction, which is perpendicular to the axis of loading

Stability of Laterally Compressed Elastic Chains

21

We tested several such samples resembling the layered geometry shown in Figure 2.1. The parts of the sheets that were either not glued or attached to the ends were uncoupled and allowed to deform freely. The gaps between the layers were introduced by inserting short strips of cardboard between the layers in the gluing zone. A schematic view of a single layer in the stack with indicated gripping, gluing and glue-free areas is shown in Figure 2.2. The details of the sample geometry and properties are given in Table 2.1. As a consequence of the sample design, the gaps between layers are very small, compared to their thickness, which constrains the deflection of the layers to be directed away from each other, producing a net repulsive effect.

4 mm

23 mm

Glue-free area

Gluing area 5 mm

Gripping area

Gripping area

25 mm

Gluing area 5 mm

4 mm

Figure 2.2. The schematic view of the sample showing gripping area (purple), gluing area (dark purple) and glue-free area (blue). For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Parameter Number of sheets Dimensions of each sheet Modulus of elasticity Poisson ratio Grip size

Value 8 25 mm × 25 mm 2 × 107 MPa (Ayrilmis et al. 2008) 0.3 25 mm × 4 mm

Table 2.1. Geometric and material properties of the cardboard samples

The compression tests were performed using a DEBEN MICROTEST tensile testing machine. The benefit of this type of machine is its extremely sensitive load

22

Modern Trends in Structural and Solid Mechanics 1

cell which allows the testing of materials and structures with little stiffness. The maximum distance between the grips of the machine is 25 mm. The load cell permits a small load resolution of 0.01 N. The samples consisted of eight 25 mm × 23 mm layers of cardboard sheets ≈1 mm thick glued along the strips ≈1 mm wide. The ends were tightly gripped by the two grips of the machine and compressed at a rate of 0.1mm/min to ensure a quasi-static loading process. The testing was carried out in a displacement-controlled mode of the machine and the load–deflection curves were recorded.

(a)

(b)

Figure 2.3. Photographs of the experimentally observed modes. (a) The top and the bottom layers exhibit two opposite deflections, while the middle layers buckle cooperatively in one direction. (b) Rapid decay of the buckling amplitude is observed along the stacking direction. For a color version of this figure, see www.iste.co.uk/ challamel/mechanics1.zip

The buckling mode shapes observed during the experiments are shown in Figure 2.3(a) and 2.3(b). Various modes were observed in the experiment, as the structure turned out to be very sensitive to any small irregularities in the layers and the coupling between them. However, a generic trend appears in most of the test cases. Either the top or the bottom layer exhibits a relatively large buckling amplitude which then reduces quickly along the stacking direction. Often, the layers in the middle would not exhibit any notable deformation, while the layers at the boundaries appear in the post-buckling regime. Even far in the post-buckling regime (as shown in Figure 2.3(a)), the layers in the central part exhibit relatively small amplitude, compared to the boundaries. This behavior resembles the classical wave localization picture (Hodges and Woodhouse 1983; Xie and Wang 1997a) but for a static problem of elastic stability. It should be noted that stacks of elastic sheets held together pose a kinematic constraint: that sheets cannot pierce each other. Unlike a chain of pendula in oscillatory motion, laterally buckled elastic sheets that are held together at the ends do not allow individual sheets to penetrate each other. This inevitably would bias

Stability of Laterally Compressed Elastic Chains

23

observed modes to be in favor of those with the highest amplitude at the edges, since the ends are free to move outwards. In a disordered pendulum, or in a model of stacked sheets that mathematically allow sheets to penetrate each other, other possibilities exist that cannot be practically realized. Motivated by the lateral instability of a stack of nominally identical elastic sheets and similarly layered materials and structures, we propose a toy problem consisting of elastically connected rigid rods that destabilize laterally beyond a critical value when compressed along the length of the rods. The rest of the chapter is concerned with this specific problem mathematically, as it provides analytical opportunities to study the localization behavior of instability modes. The simple experiment presented here is limited only as an illustration of the class of structural problems in layered media that may show similar structural behavior. 2.3. Stability of an elastically coupled cyclic chain Consider a simple chain of rigid rods coupled elastically by light springs. Although the configuration of coupled rods does not capture the complex elasticity of the layered structure, it allows analytical investigation of the effect of a regular placement of nominally identical members on the character of buckling. The effects of the disorder are treated analytically in the following sections. A chain of masses coupled with springs and coupled pendula have been traditionally used to study wave propagation (Brillouin 2003) and vibration of periodic structures in one dimension (Rosenstock and McGill 1968; Hodges and Woodhouse 1983; Castanier and Pierre 1995). Here, we propose a buckling analogy of such chains as a chain of elastically coupled rigid rods, each with one degree of freedom, as shown schematically in Figure 2.4. Each rigid rod of length L is pinned at the bottom using a hinge with a torsional spring of stiffness kt . Each rod is subjected to a compressive force P at the top acting along the length of the rods. The rods are coupled by light springs k (k  kt ), and the rotation of each rod from its equilibrium position is given by angle θ with a subscript denoting the bay number. Consider, first, the stability of an isolated rod on bearing support. The hinge allows free rotation which is resisted by a torsional spring. As the compressive loading increases from zero, the rod remains stable, while the moments about the hinge support are in equilibrium P L sin θi = kt θi . In the small-angle approximation sin θi ≈ θi , we see that there is a critical value Pcr = kt /L such that the structure is in stable equilibrium for P < Pcr and destabilizes when P = Pcr . Here, a bifurcation occurs, leading to a left or a right buckling mode characterized by the deflection angle θb when the bifurcation point is reached.

24

Modern Trends in Structural and Solid Mechanics 1

k

P

P

P

P

unit-cell P

P

P

P

k θi L 1

2

3

4

i

kt

N −2 N −1

N

Figure 2.4. A chain of N identical rigid rods. Each rod has length L and is attached to a hinge bearing and a torsional spring kt . The rods are coupled with light springs k and each is loaded with a force P . The deflected shape of each rod when buckled is indicated with dashed gray lines and the angle of deflection is labeled by θ. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Now, consider a rigid rod coupled by a light spring k to its neighbors in a cyclic chain. We designate the i-th rod in the chain, as shown within the dashed rectangle in Figure 2.4, as a unit cell within the chain. Applying the moment equilibrium condition for a unit cell leads to a recursive relation between the rotations of the neighboring rods −kL2 θm−1 + (kt + 2kL2 )θm − kL2 θm+1 = P Lθm .

[2.1]

We divide both sides of equation [2.1] by L and introduce stiffness parameter k¯0 = kt /L + 2kL and coupling parameter k¯ = kL. Then, rearranging terms, we can write ¯ m−1 + k¯0 θm − kθ ¯ m+1 = P θm −kθ

[2.2]

Now, consider a chain of N rods forming a closed loop. This implies that the rotation of the first rod is equal to that of the last θ1 = θN +1 . Equation [2.2], along with the cyclic boundary conditions, can be written in the form of an eigenvalue problem in terms of a circulant matrix ⎡¯ ⎤ ⎧ ⎫ ⎫ k0 −k¯ · · · −k¯ ⎧ θ1 ⎪ ⎪ ⎪ ⎪ θ1 ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎢ ⎪ ⎥ . ⎨ θ2 ⎪ ⎬ ⎢−k¯ k¯0 . . . .. ⎥ ⎨ θ2 ⎬ ⎢ ⎥ = P , [2.3] . . ⎢ . . . . ⎥ .. ⎪ .. ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎣ .. . . . . .. ⎦ ⎪ ⎪ ⎪ ⎪ ⎪ ⎩ ⎭ ⎩ ⎭ θN θN −k¯ · · · −k¯ k¯0

Stability of Laterally Compressed Elastic Chains

25

where the column vector on both sides is the eigenvector and P is the corresponding eigenvalue. More compactly, Cs = P s, where C is the circulant matrix of the chain. The components θm form a vector of generalized coordinates s, and the eigenvalue P is interpreted as the critical buckling load of the chain. The N eigenvectors sn , n = 1, . . . , N of a circulant matrix are well known (see, for example, Gray (2005) and Olson et al. (2014)), and have components sn,m := (sn )m that are the N -th roots mn of unity (zN )N = 1, so that sn+1,m+1 = zN :   1 2πi (n−1) 2(n−1) (N −1)(n−1) , zN , . . . , zN , zN := e N . [2.4] 1, zN sn = √ N The matrix of eigenvectors of the circulant matrix is a Fourier matrix (Olson et al. 2014) ⎤ ⎡ 1 1 1 ··· 1 N −1 2 ⎥ ⎢1 z 1 zN ··· zN ⎥ ⎢ N 2(N −1) ⎥ 1 ⎢ 2 4 1 z z · · · z ⎥ ⎢ N N N [2.5] F= √ ⎢ ⎥ .. .. N ⎢ .. .. .. ⎥ . . . ⎦ ⎣. . (N −1)(N −1) N −1 2(N −1) zN · · · zN 1 zN where each entry is the buckling amplitude sn,m = θm ; here, m indicates the location of the rod and n is an index that labels each eigenvector. To obtain the eigenvalues, we substitute θm back into equation [2.2] and obtain ¯ −kz N

n(m−1)

nm ¯ n(m+1) = Pn z nm . + k¯0 zN − kz N N

nm Dividing both sides of the above equation by zN yields

¯ −n + k¯0 − kz ¯ n = Pn . −kz N N Further, we scale the critical buckling load (eigenvalue) Pn by the critical load of a single rod Pcr = kt /L and use e−iϕ + eiϕ = 2 cos ϕ and 1 − cos ϕ = 2 sin2 (ϕ/2). Thus, the n-th scaled critical buckling load of the chain is   π(n − 1) Pn = 1 + 4 = 1 + 4 k sin2 (ϕn ), k sin2 [2.6] N where  k = kL2 /kt is the ratio of the coupling strength to rod stiffness. We have introduced ϕn = π(n − 1)/N in equation [2.6], which we call a general mode number, analogous to the reciprocal wave numbers (vectors) used in wave propagation literature (Ziman 1972). The periodicity of the eigenvalues Pn is thus analogous to those in the first Brillouin zone. Each critical buckling load is a function P (φ) = 1 + 4 k sin2 (φ) defined over the discrete set of N points ϕj , j = 1, . . . , N ,

26

Modern Trends in Structural and Solid Mechanics 1

so that Pn = P(ϕn ). Since zN live on a circle, the Brillouin zone is closed under the addition of mode number: for any two modes ϕk , ϕl , P(ϕk + ϕl ) = P (ϕk+l ) = Pn+m . All the buckling loads that correspond to the first Brillouin zone appear within a single group with the magnitudes ranging between Pmin = 1 and k, as shown in Figure 2.5 for  k = 0.1. Further, we will refer to such a Pmax = 1 + 4 packed group of buckling loads as an instability cluster. 1.4

Highest mode of the cluster

1.3

Pn

1.2 Degenerate pairs

1.1 1.0 0.9

Lowest mode of the cluster 1 N

2 N

1 2

or

N −1 N

N +1 2N

General mode number,

n−1 N

Figure 2.5. The plot of Pcr versus general mode number (n − 1)/N (blue line) for an infinite chain N → ∞. It contains the instability cluster of the arbitrarily long chain. The mode types associated with the chosen general mode numbers are shown with the subplots. For a color version of this figure, see www.iste.co.uk/ challamel/mechanics1.zip

In Figure 2.5, we plot the values of the n-th buckling load of the chain Pn versus the general mode number (n − 1)/N , using equation [2.6] for n = 1, . . . , N . The discrete points form a continuous blue line in Figure 2.5 if the matrix size N → ∞ (infinite chain). Indeed, in the limit of an infinite chain, the discrete variable (n−1)/N becomes a continuous one, which is reflected by a continuous blue curve in Figure 2.5. In the same figure, we represent results for a finite chain of size N = 10 using red dots on the blue curve. To illustrate the nature of the mode shapes, we present four examples of instability mode shapes for N = 10. In the case of a structure with one degree of freedom, the first instability cluster plotted in Figure 2.5 is the entire spectrum of the eigenvalue problem. The cluster is symmetric with respect to n = N/2 for even N N and n = (N + 1)/2 for odd

Stability of Laterally Compressed Elastic Chains

27

N . The exceptions to this rule are the lowest (n = 1) and the highest (n = N/2 or n = (N + 1)/2) buckling loads. This implies that the cluster contains pairs of degenerate buckling loads. Consider the selected mode shapes plotted in Figure 2.5. The smallest buckling load corresponds to the mode with the least energy – all rods deflect in the same direction, and coupling springs do not contribute to the strain energy. This is the buckling mode practically observable in the chain. The highest buckling load of the cluster is associated with the mode in which all the coupling springs contribute equal amounts to the energy. Any modes associated with the degenerate pairs are mere left–right reflections since the chain is invariant under the transformation (θm ) → −(θm ) for all m. The modes suggest that the deflection of the m-th rod in the chain is governed by the spatial modulation of the n-th eigenvector of matrix C. 0.15

Extended mode, n=3

2 nodes

θm 0.00

-0.15

0.15

Extended mode, n = 2

0

10

20 30 40 Bay number m

2 nodes

0.00

50

-0.15

0

10

20 30 40 Bay number m

50

Figure 2.6. The angular displacement θm of the m-th rod plotted versus its number m. Extended buckling mode 2 (left) and extend buckling mode 3 (right) appear as spatially modulated values of θm . The spatial coordinate is discrete and corresponds to rod numbers m. For a color version of this figure, see www.iste.co.uk/ challamel/mechanics1.zip

For example, we plot the spatial modes 2 and 3 in Figure 2.6 where each blue circle is the buckling angle of the m-th rod in the chain θm , plotted as a function of its position m. Since the chain is discrete, m is a discrete representation of the continuous spatial variable x along the chain. We see from the figure that the rod amplitude is modulated by a periodic function. Thus, we observe here a buckling counterpart of Bloch waves in the one-dimensional periodic chain. Note the key aspects of the behavior specific to this problem. In contrast to wave propagation and vibration, here we are studying a static problem. Hence, the chain assumes a static configuration such that the buckling amplitudes of the rods are modulated in space due to the periodicity of the structure. Here and later, we call such spatial modulation of buckling amplitudes of discrete members an extended buckling mode. This name is chosen by analogy with the wave propagation literature

28

Modern Trends in Structural and Solid Mechanics 1

where an extended mode is spatially spread and is associated with the conductivity of matter (Ziman 1972). Note that only the lowest mode of the chain of rigid rods is achievable, in practice, as the chain fails above the lowest buckling load. To observe the second mode, we need to fix a single rod on the chain to provide an artificial node to the extended buckling mode. Similarly, higher modes are achieved by providing more nodes. 2.4. Elastic stability of two coupled rods with disorder Until now, we studied a case of perfect periodicity – a cyclic chain of identical rigid rods. However, as any real structure would inevitably possess some degree of disorder, we need to account for it in our modeling. Coupled pendula serve as excellent models in structural vibration and elastic waves to study the dependence of eigenvalues of system parameters analytically. Here, we analytically consider the stability of a pair of elastically coupled rods under axial loading with the disorder. Such a model allows us to derive the exact dependence of the critical buckling loads on the structural parameters and the irregularity. P P k

L + ΔL

θ1

kt

L − ΔL

θ2

kt

Figure 2.7. A model of two rigid rods on hinge supports with torsional springs kt coupled by a weak spring k and loaded with a compressive loading P . The rods have a length imperfection of ±ΔL. For a color version of this figure, see www.iste. co.uk/challamel/mechanics1.zip

Here, we start by considering two rigid rods coupled by a spring and assume that they possess a small irregularity. For example, let their lengths be slightly different, as shown in the schematic diagram in Figure 2.7. Moment equilibrium about hinges under the small-angle approximation results in a system of two linear equations  P L1 θ1 = kt θ1 + k [L1 θ1 − L2 θ2 ] L1 [2.7] P L2 θ2 = kt θ2 + k [L2 θ2 − L1 θ1 ] L2 ,

Stability of Laterally Compressed Elastic Chains

29

where L1 = L + ΔL and L2 = L − ΔL. We divide the first equation by L + ΔL and the second equation by L − ΔL and write ⎧ kt ⎪ ⎨−P θ1 + θ1 + k [(L + ΔL)θ1 − (L − ΔL)θ2 ] = 0, L + ΔL [2.8] k t ⎪−P θ + ⎩ θ2 + k [(L − ΔL)θ2 − (L + ΔL)θ1 ] = 0, 2 L − ΔL which can be arranged formally as the following generalized eigenvalue problem: Kθ = P Kg θ.

[2.9]

Here, K is the stiffness matrix, Kg is the geometric stiffness matrix and θ = [θ1 , θ2 ] contain the buckling angle amplitudes of each rod. Note that due to normalization by L − ΔL and L + ΔL, the geometric stiffness matrix is the identity matrix. The expressions for the two eigenvalues P1,2 are easy to calculate analytically as P1,2 =

kt L2 + kL2 (L2 − ΔL2 ) kt (L2 − ΔL2 )  L kt2 ΔL2 + k(L2 − ΔL2 ) (kL4 − ΔL2 (kL2 + 2kt )) . ∓ kt (L2 − ΔL2 )

[2.10]

Let us introduce a non-dimensional measure of disorder ε as the ratio between the perturbation in rod length ΔL and the nominal rod length L, i.e. ε = ΔL/L and the 2 ¯ dimensionless coupling  k = kL/k t = kL /kt as a ratio of the coupling stiffness k and the torsional stiffness of each rod kt . We then define a dimensionless buckling load as P = P/Pcr , where Pcr = kt /L is the buckling load of an individual rod. Therefore, the expressions in equation [2.10] can be written in the dimensionless form      k(1 − ε2 ) ε2 1 − 2   1 k+ ∓  . [2.11] k2 + P1,2 =  1 − ε2 (1 − ε2 )2 k and disorder ε. We assume disorder Note that P1,2 depend only on the coupling  to be small, i.e. ε  1, and use a power series expansion for P1,2      P1 = 1 + 2 − 21k ε2 + O ε3 [2.12]   P2 = (1 + 2 k) + 1 ε2 + O ε3 . 2 k

Equation [2.12] provides a leading order approximation of the exact functional forms in equation [2.11] for small disorder.

30

Modern Trends in Structural and Solid Mechanics 1

2.0 (a) 1.8 1.6  P 1.4 1.2 1.0 0.8 0.6

 k = 0.01

(b)

 k = 0.05

(c)

 k = 0.2

-0.2 0.0 0.2 ε -0.2 0.0 0.2 ε -0.2 0.0 0.2 ε P1 approx P2 approx P1 exact P2 exact Figure 2.8. Plots of the first and the second critical buckling loads of two coupled rods as functions of disorder ε for non-dimensional coupling (a)  k = 0.01, (b)  k = 0.05 and k = 0.2. The solid lines represent the exact solution from equation [2.11], while the (c)  dashed lines correspond to the second-order series expansion from equation [2.12]. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

It is apparent that when ε = 0, we have a perfectly ordered structure with critical k. These are the same as the lowest and the highest loads P1 = 1 and P2 = 1 + 2 buckling loads of an ordered chain given by equation [2.6]. This is consistent with the Floquet–Bloch theorem as the solutions of the unit cell should specify the fundamental states which are then extended periodically for an entire medium. When ε is close to zero, the critical buckling loads have a functional dependence on disorder, given by equation [2.12]. As the disorder increases further, the approximate solution is expected to diverge from the exact solution in equation [2.11]. The change of the solution with the disorder ε for various values of coupling is demonstrated in Figure 2.8(a) for  k = 0.01, Figure 2.8(b) for  k = 0.05 and Figure 2.8(c) for  k = 0.2. The buckling load trajectories, as functions of the disorder parameter ε, show differences for different coupling strengths. This dependence is either magnified by small coupling or reduced by a strong coupling. The discussed model is particularly useful to demonstrate how buckling of a structure made of identical members is affected by coupling and by the disorder. We generalize it as follows. The system matrix A = K−1 g K of the two identical uncoupled structural members is   K 0 , [2.13] A= 0 K where K is some stiffness function of structural parameters. The eigenvalues of A are identical for all values of structural parameters and constitute the buckling loads of uncoupled members. We now introduce coupling between the members via a coupling matrix     k −k K 0 . [2.14] + A= −k k 0 K

Stability of Laterally Compressed Elastic Chains

31

Here, k is the coupling coefficient that may also depend on several structural parameters. Then, the eigenvalues of the system matrix are λ1 = K, λ2 = K + 2k. When plotted as functions of one or two parameters of interest, the eigenvalues appear as trajectories or surfaces in the parameter space. The eigenvalue trajectories (surfaces) coincide in the absence of coupling. The introduction of coupling then essentially breaks the degeneracy, and the trajectories (surfaces) separate. However, if k = 0 in a point (or on a line) in a parameter plane (or space), this causes the trajectories (surfaces) to cross over that point (line). Finally, let us introduce small irregularities, via a disorder matrix, to the sum of the previous matrix A as       1 0 k −k K 0 + + A= . [2.15] −k k 0 K 0 2 In this case, the crossing in the crossing point (crossing line) is removed and replaced by the repulsion of the respective trajectories (surfaces), as can be seen from Figure 2.8. Note that the trajectories as functions of the disorder are studied. By writing a generic parameter-dependent system matrix as a sum of member stiffnesses, coupling and disorder, we present a comprehensive matrix description of all structures considered in the present study. All specific problems are partial examples of this representation. The matrix A can be large and can have circulant or tridiagonal structure. The current discussion will still be valid. 2.5. Spatial localization of lateral buckling in a disordered chain of elastically coupled rigid rods Now, we extend the analysis for two rods to a long chain of coupled rods. For now, consider all rods to be identical (no disorder). We also abandon cyclic periodicity θ1 = θN +1 . Therefore, the system matrix C in equation [2.16] is tridiagonal ⎤ ⎡¯ k0 −k¯ · · · 0 ⎢ .⎥ ⎢−k¯ k¯0 . . . .. ⎥ ⎥ ⎢ [2.16] C=⎢ . . ⎥. ⎣ .. . . . . . . .. ⎦ 0 · · · −k¯ k¯0 Note that C1,N = CN,1 = 0, unlike in equation [2.3]. Here, as before, k¯ = kL and ¯ k0 = kt /L + 2kL, where k is the strength of the coupling spring and kt is the strength of the torsional spring in each rod. The exact spectrum of the tridiagonal matrix that describes the Coxeter graph Al for l ≥ 2 is known (Goodman et al. 2012). It is easy to show that the matrix C can be transformed into a form consisting of Al and the ¯ we can rewrite the matrix as identity matrix. Denoting β = k¯0 /k, ¯ l, C = βI − kA

[2.17]

32

Modern Trends in Structural and Solid Mechanics 1

where I is the identity matrix and Al is the adjacency matrix of the Al -Coxeter graph ¯ (a linear chain). The eigenvalues of C are then β, plus the eigenvalues of Al times k. Hence, after rescaling the eigenvalues P by the critical load of a single rod P/Pcr = 2 ¯ kt /L and considering that  k = kL/k t = kL /kt , the scaled buckling loads of the chain are Pn = 1 + 2 k − 2 k cos ϕn .

[2.18]

Here, we see that the exact calculation of the critical loads of an infinite but ordered chain is still possible. We now abandon the assumption that all rods are identical and proceed with introducing disorder. Note that there are no known methods to calculate the exact spectrum of matrices, such as the one in equation [2.16] but with random elements. Therefore, iterative methods are often employed. The transfer matrix method is extensively used in the literature on waves in periodic media and has long become a standard tool (Benaroya 1997). It is particularly efficient in treating perfectly ordered structures with small random irregularities (Xie and Wang 1997a, b). Here, we start with the definition of the transfer matrix for all identical rods and then introduce disorder into it. We use equation [2.2] to formulate a transfer matrix Tm , which controls how the state vector associated with the m-th rod sm = {θm , θm−1 } is changed to that corresponding to the m + 1-th rod: sm+1 = {θm+1 , θm }       θm+1 θm (2 + 1 − P ) −1 k k = . [2.19] θm θm−1 1 0 2 Here,  k = k Lkt and P = P kLt . In the vector–matrix notation

sm+1 = Tm sm ,

m = 1, . . . , N,

[2.20]

where Tm is the transfer matrix between the state of the unit cell m + 1 described by the state vector sm+1 and that of the unit cell m described by the state vector sm in a chain with N unit cells (rods). The whole chain or rigid rods may thus be described by a cumulative product of transfer matrices sN = TN −1 TN −2 · · · Tm · · · T2 s1  BN s1 .

[2.21]

Given a small disorder in each rod, either to the length of each rod L, to the spring coupling k or the torsion coefficient kt , the transfer matrices Tm become unequal, and the system of N coupled rods is no longer perfectly ordered. In what follows, we will treat the disorder as random. Anderson (1958) studied the asymptotic behavior of disordered lattices to show that the exponential decay of wave-function amplitude due to disorder leads to the

Stability of Laterally Compressed Elastic Chains

33

development of an insulating state of a conductor, as all electrons appear in highly localized states. Later, the ergodic multiplicative theorem by Oseledets (1968) and Furstenberg’s theorem on non-commuting random products (Furstenberg 1963) were developed to characterize the asymptotic behavior of products of random matrices as Lyapunov exponents. The first adaptations of these two theorems to the cases of vibration of multi-span beams were made by Ariaratnam and Xie (1995) and Castanier and Pierre (1995). At the same time, Xie (1995) introduced the idea of products of random matrices to studying multi-span beams in buckling. Here, we apply the analytical framework proposed by Xie (1995) to a different buckling problem. Note that, in the present study, the direction of periodicity is lateral to that of loading, in contrast to multi-span beams compressed axially. Furstenberg’s theorem allows the asymptotic behavior of the products in equation [2.21] to be calculated, quantifying the average change of norms under successive transformations Tm : sm → sm+1 . The symbol · below refers to a choice of norms, for both vectors, the L2 -norm and, for matrices, the operator norm, L∞ . For the case at hand, the determinants of all Tm are 1, ensuring that the transformations are volume-preserving. Under these conditions, for Tm , m = 1, . . . , N , non-singular, independent, identically distributed d × d random matrices, such that no two matrices share common eigenvectors, the following two limits exist: 1 1 ln BN s1 = lim ln TN −1 · · · Tm · · · T2 s1 = λ N →∞ N N →∞ N

[2.22]

1

BN x

ln BN = λ, where BN = max . x N →∞ N

x

[2.23]

lim

and lim

The application of Furstenberg’s theorem to the chain of rigid rods described by equation [2.21] leads to

BN s1

ψN + 1 = eλ ⇒ lim = eλ 1 N →∞ BN −1 s1

N →∞ 1 + ψN −1 lim

[2.24]

where ψj = (θj /θj−1 )2 . In the N → ∞ limit, ψN = ψ independent of N , and the solution to the resulting quadratic equation for ψ yields √ lim (θN /θN −1 ) = exp(λ), −1. [2.25] N →∞

It expresses how the angles in the configuration change from rod to rod and characterizes the exponential localization of the angles of rotation. For the entire chain, this relation can be rewritten as |θN +1 | = eλN |θ1 |.

[2.26]

34

Modern Trends in Structural and Solid Mechanics 1

If λ > 0, the angles of rotation will grow exponentially along the chain. At the same time, the extended modes of the structure are characterized by the Fourier matrix or complex exponents. If real Lyapunov exponents are non-zero, they govern how quickly buckling amplitudes of each rod deviate from the extended modes. The λ factors are called localization factors and are generic descriptors of the behavior in the disordered periodic structure. To understand the difference between the ordered system and the disordered one, we first calculate Lyapunov exponents for the case of identical rods and springs. For this ordered case (Tm = T for all m), BN is the largest singular value of BN which is the (N − 1)-th power of the largest singular values of T , since TN −1 , TN −2 , · · · , Tm , · · · , T2 are all identical and equation [2.21] is simplified to    2 + 1 − θN +1 k = θN 1

 P  k



−1 0

N −1 

 θ2 . θ1

[2.27]

The eigenvalues of T are Λ1,2 =

1 + 2 k − P ±

!

(1 − P )(4 k − P + 1) . 2 k

[2.28]

Xie (1995) uses Rayleigh’s quotient and Oseledets’ theorem (1968) to relate the  BN to Lyapunov exponents eigenvalues e1 = emax and e2 = emin of the matrix BN λ1 and λ2 as 1 ln emax , N →∞ N − 1 1 ln emin . = lim N →∞ N − 1

λ1 = λmax = lim

[2.29]

λ2 = λmin

[2.30]

Substituting eigenvalues of T given by equation [2.28] into the expressions for the Lyapunov exponents given by equations [2.29] and [2.30], we obtain: λ1 = max λ1,2 = max sgn(Λ1,2 ) ln |Λ1,2 |. 1,2

1,2

[2.31]

The functional dependencies of the largest factor λ1 on compressive loading  = kL2 /kt and for the P = P/Pcr = P L/kt for various values of coupling k perfectly ordered case are plotted in Figure 2.9. To understand the plot, consider that a transfer matrix in equation [2.27] is a function of the loading parameter P . If the loading is continuously increased from zero, the plot of λ1 as a function of P has three distinctive regions (as seen in Figure 2.9), each of which corresponds to a different physical response of the structure. These regions correspond to the structure being stable, buckling (instability cluster) and the failed structure for loads higher than those inside the instability cluster. The chain of rigid rods is stable until the first

Stability of Laterally Compressed Elastic Chains

35

buckling load of the cluster is reached (P < 1). In this region of P , λ1 is of no practical interest as buckling does not occur, whereas, within the instability cluster that corresponds to a set of values of P ∈ [1, 1 + 4 k], λ1 would reflect localization. Naturally, the real part of λ1 happens to be zero in this region of P, which corresponds to the absence of localization in a perfectly ordered structure. Finally, for the loads beyond the instability cluster, i.e. P > 1 + 4 k, localization is of no interest, since the structure has already failed.

5

k = 0.01

4 5

k = 0.0

3

k = 0.1 k = 0.15.2 k = 0

λ1 2 1 0 Stable structure

Instability clusters

Failed structure

P Figure 2.9. Plots of λ1 factors as functions of loading parameter P = P/Pcr for various coupling strengths  k. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

The magnitudes of λ1 for each of the three regions in Figure 2.9 become apparent when we study the eigenvalues of T given by equation [2.28] and λ1 given by equation [2.31]. Lyapunov exponents λ1 are real and positive for P < 1 and complex for 1 < P < 1 + 4 k due to the negative radical in equation [2.28]. Finally, λ1 is real and positive again for P > 1 + 4 k. The specific points that correspond to the buckling loads within the instability cluster such as P = 1, P = 1 + 4 k and others in-between correspond to Λ1 = 0 and λ1 = 0. We now introduce disorder into each rigid rod. Furstenberg’s theorem only tells us that such an asymptotic value λ exists to which the product of random matrices should converge. It does not specify how to calculate it. The previously used equations [2.28] and [2.31] are not applicable here, since equation [2.27] does not hold as all the transfer matrices are now different. Therefore, we have to use numerical algorithms to calculate localization factors. One such algorithm was employed by Xie (1995), which we will make use of as well.

36

Modern Trends in Structural and Solid Mechanics 1

Consider the disorder to be introduced into each rigid rod by changing its restoring torsional spring stiffness kt . The new torsional spring of each disordered rod has a stiffness kt = k¯t + kε , where k¯t is some mean value of torsional spring stiffness and kε is a small parameter which is a measure of disorder. Values of kε are taken from a uniform distribution. Note that in contrast to the previous case, where the irregularity was introduced via length imperfection ΔL, here, kε is a small parameter that has a dimension of torsional stiffness, i.e. Newton-meters per radian, and kε  k¯t . Recall k = kL2 /kt . Therefore, we can rewrite the first entry of the that P = P L/kt and  transfer matrix in equation [2.27] introducing the disorder in torsional stiffness as k¯t + kε 1 + ε P kt P P = 2 + = 2 + − − − , [2.32]   kL2 kL kL2 kL k k defining the dimensionless torsional stiffness perturbation ε = kε /kt . The m-th random transfer matrix obtained from equation [2.27] becomes    2 + 1+ε − P −1  k k T = . [2.33] 1 0 2+

Now, to obtain the value for the localization factor λ, we use a numerical procedure to evaluate the limit in equation [2.22]. We calculate the norm of the product of transfer matrices by the initial state vector BN numerically and then calculate the logarithm and take the limit. We perform such calculations for N = 10, 20, 50, 100 and 1000 rods, as we want to replace the theoretical limit of N → ∞ by a large enough number and be sure that the numerical calculation has converged. Practically, it turned out that the values of the localization factor do not change by more than 10% − 15%, but we plot λ as a function of P for N = 1000 to have smooth curves that are shown in Figure 2.10. Here, the localization factor λ is plotted for the same disorder magnitude ε = 0.1 but for various couplings. The plots in Figure 2.10 give valuable insight into the physics of localization of buckling modes. We observe that the strength of localization increases as coupling to disorder ratio  k/ε → 1. If the coupling to disorder ratio  k/ε 1, the chain exhibits weak localization and the respective λ curves approach those from the perfectly ordered case, as in Figure 2.9. The values of λ are no longer zero in the instability cluster. Instead, they also depend on  k/ε. In such a way, we observe that disorder affects the width of the instability cluster and the ordering of the buckling loads in it. In Figure 2.11, we plot the extended mode shapes of the chain with N = 100 rigid rods, numerically calculated from the eigenvalue problem Cε s = P s that describes the chain. Here, Cε is 100×100 matrix, similar to the one in equation [2.16] but with a small random disorder introduced in each element in the main diagonal. These modes are associated with the four lowest buckling loads P within an instability cluster of the chain. The individual angles of rotations θm , m = 0, . . . , 99 constitute the entries of the respective eigenvectors of Cε . In the case of no disorder (ε = 0), we observe

Stability of Laterally Compressed Elastic Chains

37

periodically extended modes, as expected from a perfectly ordered structure, as seen in Figure 2.11, the left column. However, even for a small disorder, the extended modes deviate from those given by the Fourier matrix, as seen in Figure 2.11, the center column. This case corresponds to weak localization that occurs for a large coupling-to-disorder ratio. However, as we increase the disorder until the couplingto-disorder ratio approaches one, we observe a strong or Anderson-type localization of the modes, as shown in Figure 2.11, the right column. Note that the location of the rod with maximum localized amplitude is random, but the exponential decay of the amplitude from this rod is symmetric, as indicated by both localization factors in Figure 2.10.

 k = 0.01

4

 k = 0.05

λ1 2 k =

0 0.5

1.0

 k = 0.1 k = 0.15 0. 2

1.5

2.0

2.5

3.0

P Figure 2.10. The localization factor λ plotted as a function of P for disorder magnitude ε = 0.1 and for various values of coupling  k. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

To better demonstrate how the buckling amplitude localizes, we choose several eigenvectors of Cε and, in Figure 2.12, plot the corresponding buckling modes of the chain for the case of the strong localization, i.e.  k/ε → 1, on the (natural) logarithmic scale ln θm , m = 0, . . . , 99. The linear dependence, with slope λ in the logarithmic plot, agrees with the limiting behavior derived in equation [2.26], that the magnitudes of θm and θm+1 are related by a multiplicative factor eλ . The lines in Figure 2.12 correspond to the Lyapunov factor λ plotted in Figure 2.10. For example, the linear regression model of the angles of rotation ln θm for mode 1 (green line) gives slope λ = 0.900 in the range of N ∈ [0, 20) and λ = −0.987 in the range of N ∈ (20, 99]. At the same time, the numerically computed critical buckling load for mode 1 is P = 0.8879. The localization factor predicted by Furstenberg’s theorem for this value of buckling load and for  k = 0.1, ε = 0.1 is obtained from the green line in Figure 2.10 and is equal to λ = 0.968, which matches the magnitude of λ computed numerically extremely well.

38

Modern Trends in Structural and Solid Mechanics 1

ε=0

θ 0.1 0.0 0 θ 0.1 0.0 -0.1

25 50 75 100 N

ε = 0.002

θ 0.2 0.1 0.0

0.5 0.0 0

θ 0.2

25 50 75 100 N

0

25 50 75 100 N

θ 0.1 0.0 -0.1

25 50 75 100 N

0

25

50 75 100 N

0

25

θ 0.5 0.25 0.0 75 100 0

50 75 100 N

25

50 75 100 N

θ 0.5

0.0 0

ε = 0.1

θ

0.0 0

θ 0.2

25 50 75 100 N

θ 0.5

0.0

0.0 0

θ 0.1 0.0 -0.1

25 50 75 100 N

0 θ 0.2

25 50 75 100 N

0.0 0

25 50 75 100 N

0

25 50 N

Figure 2.11. Four lowest extended buckling modes of a chain with N = 100 rigid rods for various magnitudes of disorder. Left column: four extended modes of the chain with no disorder, center column: four extended modes of the chain with small disorder and right column: four extended modes of the chain with moderate disorder. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Finally, we observe that the slopes of the modes 1 and 99 are the largest and the slopes of the modes in-between, such as mode 76, are much smaller. This is a reflection of the dependence of λ on the compressive loading P shown in Figure 2.10. Modes 1 and 99 correspond to the values of P at the boundaries of the instability cluster for which λ is relatively large, compared to mode 76 which corresponds to P inside the instability cluster. Therefore, we conclude that the localization factors accurately describe the behavior of the randomly disordered long chain of rigid rods and are a good measure for the localization of the bucking amplitude. Furthermore, we conclude that the symmetric exponential decay of the buckling amplitude from a random location on the chain is a generic phenomenon that depends only on the geometry of the structure, coupling between the members and the magnitude of the disorder.

Stability of Laterally Compressed Elastic Chains

ln θ 0 -6

39

ε = 0.1

-15

Mode 76

-24 -33 -42

λ

-51 Mode 99

-60 0

Mode 1 20

40

60

80

100

N Figure 2.12. Plots of the selected buckling modes of the disordered chain in a natural logarithm scale. The modes are plotted for the chain with the coupling strength k = 0.1 and disorder magnitude ε = 0.1 which constitutes the case of strong localization. For a color version of this figure, see www.iste.co.uk/challamel/ mechanics1.zip

2.6. Conclusion In this chapter, we considered the problem of elastic stability of a multilayer structure and that of an expository model, i.e. a chain of elastically coupled rigid rods. The latter is an abstraction that we hope will provide analytical insight into a host of instability problems concerning periodic elastic structures. While wave propagation and normal mode vibration of the structures with translation symmetry are extensively studied in the literature, only a few papers are devoted to the buckling of such structures. We provided a brief summary of the literature, drawing attention to the relevance of periodic structures buckling, which has been less extensively studied. It is noted that the buckling of layered structures is particularly interesting, which may possess translation symmetry along the line perpendicular to the application of load. We performed an illustrative experiment on a stack of elastic sheets glued at the ends to show that the buckling amplitude of each sheet localizes along the stacking direction, which is perpendicular to the axis of loading. We proposed a minimal model comprising of a chain of nominally identical elastically coupled rigid rods. We introduced the disorder in this model to analytically investigate the localization of buckling modes. The modeling in the absence of disorder showed that the magnitudes of the buckling loads of the chain are densely packed within a range that we termed an

40

Modern Trends in Structural and Solid Mechanics 1

instability cluster. The mode numbers associated with the buckling loads within the instability cluster appear inside the buckling analog of the first Brillouin zone. The buckling amplitudes of each rod are spatially modulated by the periodic functions along the chain and form the so-called extended buckling modes. This behavior is characterized by the eigenvectors of the chain, which are columns of the Fourier matrix. We modeled the effect of the disorder on buckling of two elastically coupled rigid rods next. We reported that the buckling mode of the structure is highly sensitive to the disorder for small coupling and relatively insensitive to it in the case of strong coupling. Therefore, the coupling to disorder ratio is the parameter that drives the sensitivity to the irregularities. Finally, we modeled a long chain of elastically coupled rigid rods with small random disorder introduced in each rod. We reported that, in contrast to the extended Fourier modes of the perfectly periodic case, the modes of the disordered chain exhibit localization. The exponential decay of the amplitude of the localized modes is determined by the Lyapunov exponent, a property introduced from the literature on products of random matrices, which we also call the localization factor. We used Furstenberg’s theorem of products of random matrices and numerical methods to calculate the localization factors for the chain and reported their values for systems of large numbers of rods. Similar to the case of two coupled rods, the chain exhibits a stronger localization as the coupling to disorder ratio reduces. For the coupling to disorder ratio close to one, we observe a strong or Anderson-type localization of buckling amplitudes along the chain. In this case, the buckling mode shows that only a few neighboring rods exhibit large amplitudes, the deviations of the remaining rods being exponentially suppressed. 2.7. References Anderson, P.W. (1958). Absence of diffusion in certain random lattices. Physical Review, 109(5), 1492–1505. Ariaratnam, S. and Xie, W.-C. (1995). Wave localization in randomly disordered nearly periodic long continuous beams. Journal of Sound and Vibration, 181, 7–22. Ayrilmis, N., Candan, Z., Hiziroglu, S. (2008). Physical and mechanical properties of cardboard panels made from used beverage carton with veneer overlay. Materials & Design, 29(10), 1897–1903. Benaroya, H. (1997). Waves in periodic structures with imperfections. Composites Part B: Engineering, 28(1–2), 143–152. ¨ Bloch, F. (1929). Uber die Quantenmechanik der Elektronen in Kristallgittern. Zeitschrift f¨ur Physik, 52(7), 555–600. Brillouin, L. (2003). Wave Propagation in Periodic Structures. Courier Corporation, North Chelmsford.

Stability of Laterally Compressed Elastic Chains

41

Byklum, E., Steen, E., Amdahl, J. (2004). A semi-analytical model for global buckling and postbuckling analysis of stiffened panels. Thin-Walled Structures, 42(5), 701–717. Castanier, M. and Pierre, C. (1995). Lyapunov exponents and localization phenomena in multi-coupled nearly periodic systems. Journal of Sound and Vibration, 183(3), 493–515. Chai, H., Babcock, C., Knauss, W. (1981). One dimensional modelling of failure in laminated plates by delamination buckling. International Journal of Solids and Structures, 17(11), 1069–1083. Cheng, B., Wang, J., Li, C. (2013). Compression tests and numerical analysis of perforated plates containing slotted holes in steel pylons. Thin-Walled Structures, 67, 129–143. Cho, M. and Kim, J.-S. (2001). Higher-order zig-zag theory for laminated composites with multiple delaminations. Journal of Applied Mechanics, 68(6), 869. Cuan-Urquizo, E. and Bhaskar, A. (2018). Flexural elasticity of woodpile lattice beams. European Journal of Mechanics – A/Solids, 67, 187–199. Dodwell, T.J. and Hunt, G.W. (2014). Periodic void formation in chevron folds. Mathematical Geosciences, 46(8), 1011–1028. Dodwell, T.J., Peletier, M.A., Budd, C.J., Hunt, G.W. (2012). Self-similar voiding solutions of a single layered model of folding rocks. Journal of Applied Mathematics, 72(1), 444–463. Elishakoff, I., Li, Y., Starnes Jr., J.H. (1995). Buckling mode localization in elastic plates due to misplacement in the stiffener location. Chaos, Solitons & Fractals, 5(8), 1517–1531. Elishakoff, I., Li, Y., Starnes Jr., J.H. (1997). Passive control of buckling deformation via Anderson localization phenomenon. Chaos, Solitons and Fractals, 8(1), 59–75. Floquet, G. (1883). Sur les e´ quations diff´erentielles lin´eaires a` coefficients p´eriodiques. Annales ´ scientifiques de l’Ecole normale sup´erieure, 12, 47–88. Furstenberg, H. (1963). Noncommuting random products. Transactions of the American Mathematical Society, 108(3), 377–428. Goodman, F.M., de la Harpe, P., Jones, V.F.R. (2012). Coxeter Graphs and Towers of Algebras. Springer Science & Business Media, Berlin. Gray, R.M. (2005). Toeplitz and circulant matrices: A review. Foundations and Trends® in Communications and Information Theory, 2(3), 155–239. Hodges, C.H. and Woodhouse, J. (1983). Vibration isolation from irregularity in a nearly periodic structure: Theory and measurements. The Journal of the Acoustical Society of America, 74(3), 894–905. Hunt, G.W., Edmunds, R., Budd, C.J., Cosgrove, J.W. (2006). Serial parallel folding with friction: A primitive model using cubic B-splines. Journal of Structural Geology, 28(3), 444–455. Juh´asz, Z. and Szekr´enyes, A. (2017). The effect of delamination on the critical buckling force of composite plates: Experiment and simulation. Composite Structures, 168, 456–464. Li, Y., Elishakoff, I., Starnes Jr., J.H. (1995). Buckling mode localization in a multi-span periodic structure with a disorder in a single span. Chaos, Solitons & Fractals, 5(6), 955–969. Li, F.M., Wang, Y.S., Hu, C., Huang, W.H. (2005). Localization of elastic waves in periodic rib-stiffened rectangular plates under axial compressive load. Journal of Sound and Vibration, 281(1–2), 261–273.

42

Modern Trends in Structural and Solid Mechanics 1

Moen, C.D. and Schafer, B.W. (2009a). Elastic buckling of cold-formed steel columns and beams with holes. Engineering Structures, 31(12), 2812–2824. Moen, C.D. and Schafer, B.W. (2009b). Elastic buckling of thin plates with holes in compression or bending. Thin-Walled Structures, 47(12), 1597–1607. Nguyen-Thoi, T., Bui-Xuan, T., Phung-Van, P., Nguyen-Xuan, H., Ngo-Thanh, P. (2013). Static, free vibration and buckling analyses of stiffened plates by CS-FEM-DSG3 using triangular elements. Computers & Structures, 125, 100–113. Olson, B.J., Shaw, S.W., Shi, C., Pierre, C., Parker, R.G. (2014). Circulant matrices and their application to vibration analysis. Applied Mechanics Reviews, 66(4), 040803. Oseledets, V.I. (1968). A multiplicative ergodic theorem. Characteristic Ljapunov, exponents of dynamical systems. Transactions of the Moscow Mathematical Society, 19, 33. Peng, L., Liew, K., Kitipornchai, S. (2006). Buckling and free vibration analyses of stiffened plates using the FSDT mesh-free method. Journal of Sound and Vibration, 289(3), 421–449. Phani, A.S., Woodhouse, J., Fleck, N.A. (2006). Wave propagation in two-dimensional periodic lattices. The Journal of the Acoustical Society of America, 119(4), 1995–2005. Pierre, C. and Plaut, R.H. (1989). Curve veering and mode localization in a buckling problem. ZAMP Journal of Applied Mathematics and Physics, 40(5), 758–761. Rikards, R., Chate, A., Ozolinsh, O. (2001). Analysis for buckling and vibrations of composite stiffened shells and plates. Composite Structures, 51(4), 361–370. Rodman, U., Saje, M., Planinc, I., Zupan, D. (2008). Exact buckling analysis of composite elastic columns including multiple delamination and transverse shear. Engineering Structures, 30(6), 1500–1514. Rosenstock, H.B. and McGill, R.E. (1968). Vibrations of disordered solids. Physical Review, 176(3), 1004–1014. Xie, W.-C. (1995). Buckling mode localization in randomly disordered multispan continuous beams. AIAA Journal, 33(6), 1142–1149. Xie, W.-C. (1997). Buckling mode localization in nonhomogeneous beams on elastic foundations. Chaos, Solitons and Fractals, 8(3), 411–431. Xie, W.-C. (1998). Buckling mode localization in rib-stiffened plates with randomly misplaced stiffeners. Computers & Structures, 67(1–3), 175–189. Xie, W.-C. and Elishakoff, I. (2000). Buckling mode localization in rib-stiffened plates with misplaced stiffeners – Kantorovich approach. Chaos, Solitons & Fractals, 11(10), 1559–1574. Xie, W.-C. and Ibrahim, A. (2000). Buckling mode localization in rib-stiffened plates with misplaced stiffeners – a finite strip approach. Chaos, Solitons & Fractals, 11(10), 1543–1558. Xie, W.-C. and Wang, X. (1997a). Vibration mode localization in one-dimensional systems. AIAA Journal, 35(10), 1645–1652. Xie, W.-C. and Wang, X. (1997b). Vibration mode localization in two-dimensional systems. AIAA Journal, 35(10), 1653–1659. Ziman, L. (1972). Principles of the Theory of Solids, 2nd edition, Cambridge University Press, Cambridge.

3 Analysis of a Beck’s Column over Fractional-Order Restraints via Extended Routh–Hurwitz Theorem

3.1. Introduction Column stability resting on elastic and viscoelastic foundations, in the presence of follower agencies, has been studied extensively (Beck 1952; Ziegler 1952; Plaut and Infante 1970; Elishakoff 2001, 2005). The fascinating dynamic stability of a column with inherent paradoxes (Smith and Herrmann 1972) depending on the kind of external restraint (Hermann–Schmidt and/or Ziegler) has attracted several scientists to help represent the mechanics of rods and columns in different engineering fields. Indeed, problems dealing with rocket thrusts, wind-exposed structures, biomechanics of the spine, as well as polymer-reinforced stratified structural glass are often modeled as structural elements subjected to follower forces. In the presence of external restraints with energy dissipating properties, different kinds of mathematical descriptions of the external restraints have been proposed. In recent years, the behavior of dissipative materials has been studied by means of fractional differential calculus in terms of the so-called Caputo fractional derivative and the Riemann–Liouville fractional integral. Indeed, the experimental data on hereditary materials show that power laws may be efficiently used to capture the material behavior (Bologna et al. 2019, 2020) and, in a rheological context (Mainardi 2010), it leads to the introduction of the springpot element. Other studies to capture the material behavior in stochastic dynamics have been proposed recently (Spanos

Chapter written by Emanuela B OLOGNA, Mario D I PAOLA and Massimiliano Z INGALES. Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

44

Modern Trends in Structural and Solid Mechanics 1

and Miller 1994; Pinnola 2016). Springpot is a one-dimensional element defined in terms of two parameters, i.e. the order of the springpot β, 0 ≤ β ≤ 1 and the FTβ anomalous modulus Cβ > 0 with physical dimensions [Cβ ] = , which may be L2 related to the elastic properties introducing the relaxation time τ0 as Cβ = E0 (τ0 )β . Such an element with intermediate behavior among elastic springs and viscous dashpot is widely used nowadays to define several types of materials including, as limiting cases, elastic (β = 0) and viscous elements (β = 1). Moreover, a simple dβ f d0 f spring corresponds to β = 0 and = = f , while the case of β = 1 dtβ dt0 β df d f = f˙, which is a Newtonian corresponds to a first-order derivative, i.e. β = dt dt dashpot. The presence of springpot as external restraints yields, on one hand, a more reliable description of the external material foundation. On the other hand, the mathematical prediction of the dynamic stability of Beck’s column became more challenging and several studies have been devoted to this topic in recent years. A different approach of the extension of the Routh–Hurwitz criterion has been recently reported for a bed of independent springpot (Bologna et al. 2017). In this study for the fractional-order dynamical systems, the authors introduce a more general representation of the external restraint by means of the fractional Zener model. This specific restraint is a generalization of the well-known Zener models that involves a fractional Kelvin–Voigt model, complete with an elastic spring forming a series element. The presence of the Zener element represents a specific generalization of the Routh–Hurwitz theorem since, for fractional-order dynamical systems, the conventional condition involving the sign of the real part of the problem eigenvalues no longer holds. The well-known Routh–Hurwitz criterion used to handle the sign of the eigenvalues can no longer be applied. In this regard, the authors introduced a novel approach that involves a conformal mapping of the complex plane yielding complex problem eigenvalues that can be solved by the extended Routh–Hurwitz method (Bologna et al. 2017; Bologna and Zingales 2018). Some numerical examples reporting the critical loads applied to the column have been reported in the paper. 3.2. Material hereditariness Conventional approaches to material behavior has led to two limit cases: i) in the presence of a complete recovery in cycle tensile tests, the concept of material elasticity is introduced, and ii) in the presence of mechanical behavior, so that complete energy dissipation (no recovery) is observed, the concept of viscous material is reported. Real types of materials show intermediate behavior and the term

Analysis of a Beck’s Column over Fractional-Order Restraints

45

material hereditariness is reported. The term hereditariness indicates the behavior of an intermediate material between elastic solid and viscous liquid; this behavior is typical of polymers, human bones, various mortars and resins used in construction, some families of rocks and other materials. The hereditariness material is therefore characterized by two asymptotic behaviors, that of the solid elastic and that of the viscous liquid.

Figure 3.1. Creep test

Two main mechanical tests allow us to construct the hereditary behavior of materials, namely the creep and relaxation tests. In a creep test, the specimen of the material undergoes a step stress history (Figure 3.1), in which the stress is instantaneously increased to some value σ0 at t = 0 and is then fixed. The typical strain response consists of i) an instantaneous increase in strain at t = 0, followed by ii) continued straining in time at a non-constant rate and (iii) an asymptotic approach to some limit value as time increases. Let J(t, σ0 ) denote the strain at time t when the value of the stress is σ0 . Then, J(t, σ0 ) = 0 when t < 0; ii) jumps to value J(0, σ0 ) at t = 0 and iii) J(t, σ0 ) monotonically increases to the limit value denoted by J(∞, σ0 ) as t → ∞. The jump in strain J(0, σ0 ) at t = 0 indicates instantaneous elasticity. The relations σ0 versus J(0, σ0 ) and σ0 versus J(∞, σ0 ) describe, respectively, instantaneous elastic response and the long-time or equilibrium elastic response. The function J(t, σ0 ) has a different dependence on time t and stress σ0 for each material and is therefore considered to be material dependent. As we assume that a linear dependence of the step stress is involved, the J(t, σ0 ) = σ0 J(t) and the time-dependent function J(t) is known as the material creep function. In the relaxation test, the specimen is subjected to a step strain history (Figure 3.2), in which the strain is instantaneously increased to some value ε0 at t = 0 and is then fixed. The typical stress history required to produce this strain history consists of i) an instantaneous increase in stress at t = 0, followed by ii) a gradual monotonic decrease of stress at a non-constant rate and iii) an asymptotic approach to some non-zero limit value as time increases. G(t, ε0 ) denotes the stress at time t, when the value of the strain is fixed at ε0 . Then, i) G(t, ε0 ) = 0 when t < 0, ii) G(t, ε0 ) jumps to the

46

Modern Trends in Structural and Solid Mechanics 1

value G(0, ε0 ) at t = 0 and iii) G(t, ε0 ) monotonically decreases to the non-zero limit value denoted by G(∞, ε0 ) as t → ∞. The jump in stress G(t, ε0 ) at t = 0 is another indication of instantaneous elasticity. The relations G(0, ε0 ) versus ε0 and G(∞, ε0 ) versus ε0 also describe, respectively, instantaneous elastic response and the long-time or equilibrium elastic response. G(t, ε0 ) has a different dependence on time t and strain ε0 for each material and is therefore considered a material property. As the linearity condition is fulfilled, G(t, ε0 ) = ε0 G(t) and the time-dependent material function G(t) is known as a relaxation function.

Figure 3.2. Relaxation test

A generic stress/strain history, namely σ(τ ) and ε(τ ) with τ ≤ t, yields:  σ(t) =  ε(t) =



t 0

0

G(t − τ )dε(τ ) + ε0 G(t) = 

t

J(t − τ )dσ(τ ) + σ0 J(t) =

t

0

0

G(t − τ )ε(τ ˙ )dτ + ε0 G(t) [3.1a]

t

J(t − τ )σ(τ ˙ )dτ + σ0 J(t) [3.1b]

Equations [3.5a] and [3.5b] are defined in terms of the Boltzmann superposition d with dσ = σdt ˙ and dε = εdt ˙ increments, where [·] = dt . The Laplace transform of equations [3.1a] and [3.1b] with σ0 = 0 and ε0 = 0 yields the fundamental relation between the creep and relaxation of linear hereditariness in the Laplace domain as: 1 ˆ G(s) ˆ J(s) = 2 s

[3.2]

ˆ ˆ where G(s) and J(s) are the Laplace transform of relaxation and creep functions, respectively. This fundamental relationship expresses the circumstance that G(t) and J(t) are functions related to each other in the domain of Laplace. It follows that if it is determined through experimental tests G(t), the function J(t) is determined accordingly and vice versa. Another approach used to develop constitutive equations for linear hereditary response involves mechanical analogs. These are mechanical devices formed by combining linear elastic springs and linear viscous dampers in series or parallel. The

Analysis of a Beck’s Column over Fractional-Order Restraints

47

devices can be shown to exhibit a time-dependent response that is similar to that observed in hereditary materials, namely creep under a constant load and force relaxation under a constant deformation. For this reason, these devices are treated as mechanical analogs of hereditariness response. Since the springs and dampers are described by linear equations, as are the equations for the kinematics of deformation and force transmission, there is a linear relation between the overall force and deformation. These approaches are known as rheological models, which will be discussed in the following.

Figure 3.3. Classic viscoelastic model

The spring shown in Figure 3.3(a) is the elastic (or storage) element, as the force is proportional to the extension; it represents a perfect elastic body obeying the Hooke law. This model is thus referred to as the Hooke model. We denote the pertinent elastic modulus by m. In this case, there is no creep or relaxation, so the creep compliance and the relaxation modulus are constant functions J(t) ≡ 1/E, G(t) ≡ E. The dashpot in Figure 3.3(b) is the viscous (or dissipative) element, with the force being proportional to the rate of extension; it represents a perfectly viscous body obeying the Newton law. By denoting the pertinent viscosity coefficient by ν, we observe that the Hooke and Newton models represent the limiting cases of viscoelastic bodies. A branch constituted by a spring parallel to a dashpot is known as the Voigt model (see Figure 3.3(c)), where τε is the retardation time. A branch constituted by a spring in series with a dashpot is known as the Maxwell model (see Figure 3.3(d)), where τσ is the relaxation time. The Voigt model exhibits

48

Modern Trends in Structural and Solid Mechanics 1

an exponential (reversible) strain creep but no stress relaxation, which is also referred to as the retardation element. The Maxwell model exhibits an exponential (reversible) stress relaxation and a linear (non-reversible) strain creep, which is also referred to as the relaxation element. By increasing the number of simple elements to the Kelvin–Voigt model, other more accurate models are obtained in the simulation of viscoelastic behavior. Such models are called the SLS (Standard Linear Solid) or Zener model. The simplest viscoelastic model is obtained by adding a spring, either in series to a Voigt model, or parallel to a Maxwell model. In this way, according to the combination rule, we add a positive constant, both to the Voigt-like creep compliance and to the Maxwell-like relaxation modulus, so that Jg > 0 and Ge > 0. Such a model was considered by Zener with the denomination SLS, which will also be referred to here as the Zener model. In Figure 3.4(a), we have a Zener model with a spring in series to a Kelvin Voigt model.

Figure 3.4. a) Zener model: a spring in series to a Voigt model and b) Zener model: a spring in parallel to a Maxwell model

The model shown in Figure 3.4(b), consists of a Hooke model in parallel to a Maxwell model, with α and β the same for both. Similarly, the functional stress–strain relations reported have an equivalent differential formulation in terms of elastic (Hookean) and viscous (Newtonian) elements. 3.2.1. Linear hereditariness: fractional-order models Linear hereditariness is certainly the field of the most extensive applications of fractional calculus, in view of its ability to model hereditary phenomena with long

Analysis of a Beck’s Column over Fractional-Order Restraints

49

memory. The analysis starts from the power-law creep to justify the introduction of the operators of fractional calculus into the stress–strain relationship. Let us consider the hereditariness of a material with creep compliance, J(t) =

1 G0 Γ(1 + β)



t τs

β [3.3]

where Γ(·) is the Euler–Gamma function, β, and [τs ] = T are the material parameters that may be estimated through a best-fitting procedure of experimental data and G0 is the elastic modulus of the material. Such creep behavior is found to be of great interest in a number of creep experiments; it is usually referred to as the power-law creep. Using the reciprocity relationship [3.2] in the Laplace domain, we can find for such a hereditariness solid, its relaxation modulus and then the corresponding relaxation spectrum. After simple manipulations, we obtain G(t) =

G0 Γ(1 − β)



t τs

−β [3.4]

For our hereditariness solid exhibiting power-law creep, the stress/strain relationship in the creep representation can be easily obtained by inserting the creep law equation [3.3] into the integral equation [3.2]. Straightforward manipulations show that the power-law functional class in equations [3.3] and [3.4] satisfies the conjugation relation, which yields, upon substitution into equations [3.1a] and [3.1b], the following constitutive relations:  t   G0 τsβ σ(t) = (t − τ )−β ε(τ ˙ )dτ = G0 τsβ D0β+ ε (t) [3.5a] Γ(1 − β) 0  t 1  β  1 β I0+ σ (t) (t − τ ) σ(τ ˙ )dτ = [3.5b] ε(t) = G0 τsβ Γ(β + 1) 0 τsβ G0   where C D0+ β is the Fractional Caputo derivative and [Ioβ+ ] is the Riemann–Liouville fractional integral. The constitutive equations [3.5a] and [3.5b] have been modeled with the introduction of a new rheological element, the springpot (Figure 3.5), after Scott-Blair. Springpot is a mechanical element with mechanical properties that are intermediate between those of the Hooke model and a pure viscous fluid (Newton model). The use of fractional calculus in linear hereditariness leads us to generalize the classical mechanical models, in that the basic Newton element (dashpot) is substituted by springpot. The springpot (Figure 3.5) is defined in terms of two parameters, i.e. Cβ = G0 τsβ ≥ 0 and β, with β ∈ [0, 1], whose constitutive relation is given in equations [3.5a] and [3.5b]. Such an element is widely used nowadays to define several

50

Modern Trends in Structural and Solid Mechanics 1

types of materials including as limiting cases, elastic (β = 0) and viscous elements d0 f dβ f (β = 1). More precisely, a simple spring corresponds to β = 0 and β = 0 = f , dt dt df dβ f = f˙, while the case of β = 1 corresponds to a first-order derivative, i.e. β = dt dt which is a Newtonian dashpot.

Figure 3.5. Springpot element

As the springpot element is connected with springs in series and parallel, the generalization of the rheological models, as shown in section 3.2.1, is obtained (see Figure 3.6).

Figure 3.6. Fractional models

Moreover, the elastic spring and the springpot in series in Figure 3.6(c) are described by a constitutive equation σ(t) = Eε(t) + Eβ (Dβ ε)(t)

[3.6]

Analysis of a Beck’s Column over Fractional-Order Restraints

51

that is the generalization of the Kelvin–Voigt element. In the presence of spring in parallel, the constitutive equation can be written as τs (Dβ σ) (t) + σ(t) = Eβ (Dβ ε)(t)

[3.7]

with Eβ = Eτsβ and the combination corresponds to the Zener element, as shown in Figure 3.6(a), with the governing equation in the form Eβ (Dβ ε)(t) + E2 ε =

Eβ β (D σ)(t) + (1 + α)σ(t) E1

[3.8]

and as Eβ = E2 τsβ we get the constitutive equation of the fractional Zener model, in the form: τsβ (Dβ ε)(t) + ε = ατsβ (Dβ σ)(t) + γ

[3.9]

E1 + E2 K2 and α = that is formally analogous to the classical Zener E1 E2 K1 model, as we replace fractional derivative with integer-order derivative and viscosity η = E2 τ0 .

with γ =

In the next section, the stability of the column over generalized foundation is considered. 3.3. Dynamic equilibrium of an elastic cantilever over a fractional-order foundation Let us consider a Bernoulli–Euler elastic column with cross-section A, length L and rotary inertial properties across the centroidal axis denoted as J1 . The Young modulus of elasticity has been denoted as E and the material density as ρ, both of which are assumed as homogeneous along the column axis, denoted as x3 in Figure 3.7. The column is clamped at x3 = 0, which is supported by a bed of vertical and independent springpots, connected in series and parallels with two springs to compose the fractional Zener model described in the previous section, assumed to be homogeneous along the x3 -axis. The column is subjected to a follower axial load, denoted as P, at the free end (Figure 3.7). In the following, we restrict our analysis to the onset of bifurcations from the undeformed (straight) configuration (Figure 3.7) in the (x2 − x3 ) plane, defined by the vertical displacement field w(x3 ) of the column axis. The governing equation of the column may be obtained in a classical fashion with ·· ∂2 the equilibrium of a column element (Figure 3.8), where we denote [•] = 2 and Ff ∂t the reaction of the fractional-order, hereditary support of the column foundation.

52

Modern Trends in Structural and Solid Mechanics 1

Figure 3.7. Onset of the column bifurcations and its generic section

Figure 3.8. Onset of the column bifurcation

The equilibrium equation of the column can be written as (omitting time argument): N (x3 + Δx3 )cos(φ(x3 ) + Δφ(x3 )) − N (x3 ) cos(φ(x3 ))+

[3.10a]

+ T (x3 ) sin(φ(x3 )) − T (x3 + Δx3 ) sin(φ(x3 ) + Δφ(x3 )) = 0 T (x3 ) sin(φ(x3 ))+N (x3 ) sin(φ(x3 ))−T (x3 +Δx3 ) cos(φ(x3 )+Δφ(x3 ))+ [3.10b] − N (x3 + Δx3 ) sin(φ(x3 ) + Δφ(x3 )) + ρA

∂2w Δx3 + Ff (x3 ) = 0 ∂t2

T (x3 + Δx3 ) cos(φ(x3 ) + Δφ(x3 ))Δx3 + T (x3 + Δx3 ) sin(φ(x3 )+ + Δφ(x3 ))Δw(x3 ) + M (x3 ) − M (x3 + Δx3 ) = 0

[3.10c]

Analysis of a Beck’s Column over Fractional-Order Restraints

53

where, in the latter equation, we neglect the contribution of the axial stress moment that is a higher-order infinitesimal. Since the onset of bifurcation is analyzed, the following approximations hold time sin(x) ∼ = x; cos(x) ∼ = 1, yielding the equilibrium equations in linearized formulation: N (x3 + Δx3 ) − N (x3 ) + T (x3 )φ(x3 )+

[3.11a]

− T (x3 + Δx3 )(φ(x3 ) + Δφ(x3 )) = 0 T (x3 )φ(x3 ) + N (x3 )φ(x3 ) − T (x3 + Δx3 )+ − N (x3 + Δx3 )(φ(x3 ) + Δφ(x3 )) + ρA

[3.11b]

∂2w Δx3 + Ff (x3 ) = 0 ∂t2

T (x3 + Δx3 )Δx3 + T (x3 + Δx3 )(φ(x3 )+

[3.11c]

+ Δφ(x3 ))Δw(x3 ) + M (x3 ) − M (x3 + Δx3 ) = 0 After some straightforward manipulations, and neglecting higher-order contributions, the governing equations of the onset of bifurcation can be written as: ∂N (x3 , t) =0 ∂x3 ρA

[3.12a]

∂ 2 w (x3 , t) ∂T (x3 , t) ∂φ (x3 , t) + Ff (x3 , t) − − N (x3 , t) = 0 [3.12b] 2 ∂t ∂x3 ∂x3

∂M (x3 , t) = T (x3 , t) ∂x3 yielding, after substitutions φ (x3 , t) = −

[3.12c] ∂w (x3 , t) ; N (x3 , t) = −P ; M (x3 , t) = ∂x3

∂ 2 w (x3 , t) the governing equation of the transverse displacement of the column ∂x3 2 resting over a generalized foundation: −EJ

ρA

∂ 2 w(x3 , t) ∂ 4 w(x3 , t) ∂ 2 w(x3 , t) + EJ +P + s(x3 , t) = 0 4 2 ∂t ∂x3 ∂x3 2

[3.13]

where the reaction support [s (x3 , t)] = F/L, which is governed by the fractional differential equation for the fractional Zener model introduced in the previous section: Cβ Dβ w + K2 w =

Cβ β D s + (1 + α)s K1

[3.14]

where [K1 ] = [K2 ] = K/L2 are the spring stiffness of the restraint and [Cβ ] = F/L2 is the anomalous viscosity of the restraint that depends on the differentiation-order,

54

Modern Trends in Structural and Solid Mechanics 1

β, and as far as β = 1 the Newtonian viscosity η is recovered. In order to achieve a more convenient formulation for the external restraint constitutive equation, we set Cβ = K2 τsβ and, in such a context, equation [3.14] becomes τsβ Dβ w + w = ατsβ Dβ s + γ

[3.15]

The boundary value problem associated with equation [3.13] is defined by using the initial and boundary conditions: dw ∂ 2 w ∂ 3 w w (0, t) = = 0; EJ = EJ =0 [3.16] dx3 0 ∂x3 2 L ∂x3 3 L w (x3 , 0) = 0

; w˙ (x3 , 0) = 0

[3.17]

It can be observed that [3.10a], [3.10b] and [3.10c] rule the mechanics of the column resting on a fractional-order foundation, which is formally analogous to the well-known column on an elastic (Beck 1952) and viscous (Plaut and Infante 1970) foundation. The onset of stability of Beck’s column resting on the fractional-order foundation is analyzed hereinafter in the non-dimensional form obtained by casting equation [3.13] in the form: ∂2w ∂t

2

+

2 ∂4w 2∂ w + λ + τ02 s¯ (ζ, t) = 0 ∂ζ 4 ∂ζ 2

s = τsβ D0β+ w ¯+w ¯ ατsβ D0β+ s¯ + γ¯

[3.18a] [3.18b]

t w s (ζ, t¯) P AL4 P L2 x3 ¯ ; t = ; τ0 = ;w ¯ = ; λ2 = and s¯ (ζ, t¯) = L τ0 EJ L EJ k1 L with the associate boundary conditions: ∂w ¯ ¯ ¯ ∂2w ∂3w w ¯ (0, t) = (ζ, t) = (ζ, t) = (ζ, t) = 0 [3.19] 2 3 ∂ζ ∂ζ ∂ζ 0 1 1

where ζ =

Despite these formal analogies, the stability analysis of the column in the presence of fractional-order time-derivatives needs a specific core, which will be discussed in the next section. 3.4. Stability analysis of Beck’s column over fractional-order hereditary foundation In this section, the evaluation of the critical instability load of Beck’s column is outlined. Stability analysis is conducted, as usual, by introducing the variable

Analysis of a Beck’s Column over Fractional-Order Restraints

55

separation w(ζ, ¯ t¯) = v(ζ)y(t¯) and in order to satisfy equation [3.18b], analogous factorization of the load term s¯(ζ, t¯) = v(ζ)r(t¯), it is given by: d4 v (ζ) d2 v (ζ) + λ2 − ω 2 v (ζ) = 0 4 dζ dζ 2





d2 y t + τ02 r t + ω 2 y t = 0 2 dt

[3.20a] [3.20b]

with function ω 2 being an unknown separation constant. The time-dependent function r (t¯) related to the Zener foundation reaction is provided by the solution of the fractional differential equation:     ατs β D0β+ r (t¯) + γr (t¯) = τs β D0β+ y (t¯) − y(t¯) [3.21] which is coupled with the governing equations in equation [3.20a] to provide the transverse displacement of the column. The boundary conditions for the shape function v(ζ) can be written as: dv d2 v d3 v v (0) = = = =0 dζ 0 dζ 2 1 dζ 3 1

[3.22]

In addition, we observe that the governing equation for the shape function involves the parameter ω 2 that has the role of unknown dynamical frequency, as shown in equation [3.20b]. The separation constant is ω 2 ∈√ C, with C as the complex field, so that ω 2 = ωr + iωi , with ωr , ωi ∈ R and i = −1 as the imaginary unit. The generalized Zener model of external restraints reduces, for limiting values of the material parameters (i.e. α → 0, with k2 = 0) to the fractional Kelvin–Voigt model. In this case, use α → 0, with k2 → 0, so that only the springpot is retained in the external restraints. The fractional-order Maxwell model is instead obtained as α → ∞. 3.4.1. The characteristic polynomial The exact solution of equation [3.20a] involves linear combinations of transcendental and trigonometric functions that, as we introduce the boundary condition, yields the eigenvalue equation among the axial load λ and the dynamic frequency ω. Such a relation, yielding a condition ω = ω(λ), corresponds to a different time evolution of the column displacement for each choice of λ. The explicit relation ω = ω(λ) is very cumbersome, which is beyond the scope of this chapter. In the following, we resort to an approximate representation of the shape function v(ζ) by assuming a two-term linear combination as: v (ζ) = A1 φ1 (ζ) + A2 φ2 (ζ) = φj (ζ) Aj

j = 1, 2

[3.23]

56

Modern Trends in Structural and Solid Mechanics 1

where in the latter equation [3.23], we use the sum on the repeated index convention. Shape functions φj (ζ) in equation [3.23] belong to a set of functions that satisfy the dφj (ζ) = 0 and Aj are the unknown coefficients. boundary conditions φj (ζ) = dζ ζ=0 The use of the two-term approximation of the space-dependent solution v(ζ) is a well-established procedure in the context of elastic stability of non-conservative systems. The reduction method introduced in equation [3.23] yields, in the case of the purely viscous foundation obtained for β = 1 in the proposed analysis, an overestimation of the critical load of almost 4%, with respect to the exact solution of equation [3.20a]. A similar consideration holds true for the case of elastic foundation where the overestimation of the critical load obtained with the two-term expansion with respect to the exact solution is almost 1% (Plaut and Infante 1970). Introducing equation [3.23] into equation [3.20a] yields: 4

2 ∂ φj (ζ) 2 ∂ φj (ζ) 2 ε (ζ) = Aj + λ − ω A φ (ζ)

= 0 [3.24] j j ∂ζ 4 ∂ζ 2 with j = 1, 2 and the sum convention being used. The unbalance function, ε(ζ), projected onto the manifold defined by two functions φj (ζ) yields two algebraic equations in terms of constants Aj that can be written as: 

1

0



0

1



ε (ζ) φ1 (ζ) dζ = ε11 ω 2 , λ2 A1 + ε21 ω 2 , λ2 A2 = 0

[3.25a]



ε (ζ) φ2 (ζ) dζ = ε12 ω 2 , λ2 A1 + ε22 ω 2 , λ2 A2 = 0

[3.25b]

where εjk are the coefficients. The system of algebraic equation in equations [3.25a] and [3.25b] can be solved by an inhomogeneous set of constants Aj , only if the determinant:









r ω 2 , λ2 = ε11 ω 2 , λ2 ε22 ω 2 , λ2 − ε21 ω 2 , λ2 ε12 ω 2 , λ2 = 0 [3.26] that corresponds to a specific condition among the separation constant ω 2 and the square of the non-dimensional load λ. For any specified value of the axial load, a value of dynamic frequency ω 2 (λ2 ) is obtained by [3.26]. In the case of the springpot external restraint (namely α → ∞), the solution of equation [3.20b] is obtained as a series of two-parameter Mittag–Leffler functions as: y(t¯) =

∞ k  (−1) k=0

k!



ω2 τ01−β τβ

k

 (k) t2(k+1)−1 E2−β,2+βk



ω2 τ01−β τβ

 t2−β [3.27]

Analysis of a Beck’s Column over Fractional-Order Restraints

57

(k)

where Eλ,μ (·) is the two-parameter Mittag–Leffler function, with k = (0, 1, 2, . . .) (Podlubny 1998). The stability of the solution in equation [3.27] cannot be inferred by direct inspection and, in such a case, several strategies to assess the stability of dynamical systems governed by fractional differential equations (FDE) have been proposed. Moreover, the Laplace transform of equation [3.20b] yields conditions on   stability by checking the poles of the Laplace transform L y t = yˆ (s) (Atanackovic and Stankovic 2004; Atanackovic et al. 2015). Similar considerations hold true for the case of generalized Zener external restraints and, in those cases, a modal space approach leading to decoupled fractional differential equations involving the separation constant ω is obtained. In the following, a different approach to stability of FDE is proposed under the assumption of rational values of the derivation order. 3.4.2. State-space representation of the dynamic equilibrium equation The transverse displacement w(ζ, t) of cantilever columns, introduced in the previous section, is governed by the two differential equations [3.20a] and [3.20b]. Separated solutions involve the Mittag–Leffler series, which requires the knowledge of the function ω 2 = ω 2 (λ) by equation [3.20b], but no considerations on the stability of the solution may be inferred. In this section, we aim to provide a solution to the stability problem of the fractional-order Beck’s column with an original p approach. Let us assume, in the following, that the differentiation order β = , with q p, q ∈ N and p ≤ q since 0 ≤ β ≤ 1, may be expanded. Under these circumstances, the time variations of the transverse displacement y(t¯) are governed by a coupled system of FDE that can be written as (Rossikhin and Shitikova 2001): d2 y ¯ (t) + r (t¯) + ω 2 (λ) y (t¯) = 0 dt¯2

[3.28]

The composition rule of the fractional-order operators namely D0β+1 (D0β+2 f ) = allows us to recast equation [3.28] in the form:

D0β+1 +β2

2q  l=1 p  l=1

 ¯  cl D0lβ+ y (t¯) + ω 2 (λ) y (t¯) + r (t¯) = 0 p  ¯   ¯   al D0lβ+ r (t¯) + γr (t¯) − bl D0lβ+ y (t¯) + y (t¯) = 0

[3.29a]

[3.29b]

l=1

with the order β¯ = 1/q and the coefficients cl = 0 ∀l = 2q, c2q = 1, al = bl = 0∀l = p al = τsβ and bl = τsβ for l = p .

58

Modern Trends in Structural and Solid Mechanics 1

Under these circumstances, equation [3.29a] can be expanded in terms of a state variable vector x(t¯) with order mxp, with m = 2q: ⎡ ⎤ ¯  y ¯(t)  ⎥ ⎡ ⎤ ⎢ Dβ+ y ⎢ ⎥ y1 (t¯) ⎢  0¯  ⎥ 2β ⎥ ⎢ y2 (t¯) ⎥ ⎢ D0+ y ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ y3 (t¯) ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ .. ⎥ ⎢ ⎥ ⎢ .. ⎢ ⎥ ⎢  .¯  ⎥ . ⎥ ⎢ ⎥ ⎢ D0r+β y ⎥ ⎢ yp+1 (t¯) ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎢ ⎥ ⎢ ⎥ . . .. .. [3.30] x (t¯) = ⎢ ⎥ ⎥=⎢ ⎥ ⎢ ⎥ ⎢  ¯ ⎥ ⎢ ym (t¯) ⎥ ⎢ (m−1)β y ⎥ ⎢ ⎥ ⎢ D0+ ⎥ ⎢ r1 (t¯) ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ¯ ⎢ r2 (t¯) ⎥ ⎢  p((t)  ⎥ ⎥ ⎢ ⎥ ⎢ β¯ D0+ r ⎥ ⎢ ⎥ ⎢ .. ⎢ ⎥ ⎣ ⎦ . ⎢ ⎥ .. ⎢ ⎥ ¯ . rp (t) ⎣ ⎦ (p−1)β¯ r D0+ with the supplementary 2q − 1 and p − 1 identities:  ¯  D0β+ yl (t¯) − yl+1 (t¯) = 0 l = 1, 2, ..., 2q − 1  ¯  D0β+ rl (t¯) − rl+1 (t¯) = 0 l = 1, 2, ..., p − 1

[3.31a] [3.31b]

The introduction of the state variable vector and the state identities into equations ¯ [3.30] and [3.31a] allows us to express equation [3.29a] as a system of β-order FDE as:  ¯  c1 1 D0β+ y1 (t¯) + h12 y2 = 0 .. [3.32] .   ¯ cmm D0β+ ym−1 (t¯) + r (t¯) + hm1 y1 (t¯) = 0 with hi,i+1 = 1 for i = 1, 2, ..., 2q − 1 and , hm1 = ω 2 , h12 = h23 = h34 = · · · = 1. The supplementary FDE for the state variable of the Zener foundation can be written as:  ¯  cm+j,m+j D0β+ rj (t¯) + hm+j,m+j+1 rj+1 = 0 ..  ¯.   ¯  cm+p,m+p D0β+ rp−1 (t¯) − hm+p,m+1 r1 (t¯) + cm+p,m D0β+ yp−1 (t¯) +hm+p,1 y1 (t¯) = 0 [3.33]

Analysis of a Beck’s Column over Fractional-Order Restraints

59

with cij = δij and δij as the Kronecker delta for i = m and i = m + p and j = p. cmm = τ02 and cm+p,1 = ατsβ and cm + p, p = τsβ . The system of equation [3.32] can be written in the matrix form as:  ¯ 

C D0β+ x (t¯) + H ω 2 x (t¯) = 0 and in the matrix form: ⎡ 1 0 0 ··· . . ⎢ ⎢ 0 1 .. . . ⎢ ⎢ .. ⎢ . 0 1 ... C=⎢ ⎢. . . ⎢ .. .. . . 1 ⎢ ⎢ ⎣0 0 0 ··· 0 0 · · · τsβ

0 ..

.

..

.

..

. ···

0

[3.34]



⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ 0 ⎦ ατsβ 0 .. . .. .

[3.35]

and the frequency-dependent matrix H ω 2 is given by: ⎡

⎤ 0 0 ⎥ ⎥ ⎥ ···⎥ ⎥ ⎥ ···⎥ ⎥ 0 ⎥ ⎥ ⎥ 1 ⎦ 1 0 · · · · · · 0 αγ 0

01 ⎢0 0 ⎢ ⎢ .. .. ⎢. . ⎢ ⎢ H(ω 2 ) = ⎢ ... ... ⎢ ⎢0 0 ⎢ ⎢ . ⎣ 0 ..

··· ··· ··· ··· ··· ··· .. . ··· ··· .. ··· . ··· ··· · · · ω2 1 · · · .. . . · · · .. · · · 0 1 .. .

Equation [3.34] can be rewritten in the form:  ¯  D0β+ x (t¯) + V(ω 2 )x (t¯) = 0

[3.36]

[3.37]

matrix of the system related to the with matrix V ω 2 the dynamic

frequency-dependent matrix H ω 2 as: V(ω 2 ) = C−1 H(ω 2 )

[3.38]

The system of fractional differential equations given in equation [3.37] is stable as the eigenvalues of the dynamic matrix lie in a specific sector of the complex plane. In order to investigate system stability, we find a solution in the form:   ¯ x (t¯) = ΦEβ −t¯−β χ [3.39]

60

Modern Trends in Structural and Solid Mechanics 1

with Φ ∈ Cm and χ ∈ C are the eigenvectors and eigenvalues of the matrix C(ω 2 ). The substitution of equation [3.39] into equation [3.34] yields the algebraic system: (Iχ − C(ω 2 ))Φ = 0 where I is the identity matrix to the eigenvalue equation:

det[C ω 2 − Iχ] = 0

[3.40]

[3.41]

Equation [3.40] is an algebraic equation for the problem eigenvalues χ(ω 2 ), which can always be expressed as: χm+1 + χm−1 Cβ + Am (ω 2 ) = 0

[3.42]

where Am is a β-dependent coefficient that will be specified as the order of fractional derivative is prescribed. The explicit values of the roots χ1 (ω), · · · , χm (ω) of the algebraic secular equation in equation [3.42] cannot be obtained for any values of the order β, and some explicit values have been obtained in previous papers (Bologna et al. 2017). In the following, we introduce a different approach to the stability of fractional-order Beck’s column based on the extension of the Routh–Hurwitz criterion. As far as the assumption of rational order derivative is removed, then the real (or irrational) order of derivation may be decomposed, as an example, in terms of the series of partial fractions and an unbounded number of state functions will be involved in the analysis. 3.4.3. Stability analysis of fractional-order Beck’s column via the extended Routh–Hurwitz criterion The direct solution of the secular equation in equation [3.42] provides the values χj (ω 2 ) = χj (ω 2 (λ)) j = 1, 2..., m of the system eigenvalues that can be χj ∈ R or χj , χ∗j ∈ C, where ∗ denotes the complex conjugate. The dynamic stability of the fractional-order set of differential equations in equation [3.32] is provided as (Marden 1949; Matignon 1996; Barnett 1970): π β¯ arg (χj ) ≤ ∀j = 1, 2, . . . , m [3.43] 2 where we denote arg(χj ) as the argument of the complex  number  χj evaluated in the −1 χj,I , with χj,I and χj,R principal Riemannian manifold that is arg(χj ) = tg χj,R as the imaginary and real components of the complex number χj , respectively. Stability analysis conducted by the direct evaluation of the roots of the characteristic equation [3.42] is a formidable task since no closed-form expression of

Analysis of a Beck’s Column over Fractional-Order Restraints

61

the roots χ(ω 2 ) is available under the general differential order β. The search for eigenvalues χj (ω 2 ) that fall into the instability region (Figure 3.9) can be observed under cases β = 1, resorting to the well-established Routh–Hurwitz criterion arg(χj ) ≤ π2 ∀j, which corresponds to χj.R ≤ 0 ∀j (Marden 1949). Similar considerations also hold for the case β = 0, yielding dynamic stability as χj ∈ R− ∀j. In addition, we observe that under such conditions, the Hermann–Smith paradox is involved, meaning that the stability load λ is not affected by the external elastic restraints (β = 0) (Beck 1952). Indeed, it is well known that the evaluation of the number of roots χj j = 1, 2 . . . m with positive real parts is provided by the evaluation of the number of change of signs of the determinant sequence (Marden 1949) Δ1 , Δ2 , . . . Δn , namely V (1, Δ1 , Δ2 , . . . , Δn ), with V being the number of signs changed. The Routh–Hurwitz theorem in (Marden 1949) can be obtained by the Sturm sequence applying the characteristic equation, followed by an appropriate rotation of the complex plane counterclockwise by the phase θ = iπ/2. A different case is involved as fractional-order hereditary restraints are considered with 0 ≤ β¯ ≤ 1, as shown in equation [3.29a]. Indeed, under such circumstances, a vector of the complex phase with origin in x = 0 corresponds to dynamic instability ¯ ¯ βπ βπ as − ≤ arg(χj ) ≤ , as shown in Figure 3.9. 2 2

Figure 3.9. Stability region on the complex plane

In order to extend the Routh–Hurwitz stability theorem to this latter case, we consider the complex of conjugated χj roots on the borders of the sector with arguments: arg(χj ) =

¯ βπ 2

arg(χ∗j ) = −

¯ βπ 2

[3.44]

62

Modern Trends in Structural and Solid Mechanics 1

The introduction of phase shifts ϕj = (1 − β)

π 2

ϕ∗j = (1 + β)

π 2

[3.45]

as: χ ¯j = χj eiϕj ;

χ ¯∗j = χ∗j eiϕj



[3.46]

yields arguments of the couples χ ¯j as: arg(χ ¯j ) =

π ; 2

arg(χ ¯∗j ) = −

π 2

[3.47]

which corresponds to complex roots on the imaginary axis. Complex conjugated ¯ βπ roots in the instability region with arg(χj ) ≤ yield after the phase shifts, 2 Re(χ ¯j ) > 0; Re(χ ¯∗j ) > 0. Complex conjugated roots within the stable region with βπ ∗ arg(χj ) ≥ βπ ¯j ) ≤ 0; 2 and arg(χj ) ≥ 2 yield, after the phase shifts: Re(χ ∗ Re(χ ¯j ) ≤ 0. Under these circumstances, as we introduce the phase shifts in equation [3.45] as: χj = χ ¯j e−iϕj

[3.48]

and we replace this in equation [3.42], we obtain a complex coefficient eigenvalue equation in terms of roots χ ¯j : s(ω 2 ) = Nm+1 χ ¯m+1 + Nm−1 χ ¯m−1 Cβ + Am (ω 2 ) = 0

[3.49]

where Nm+1 = e−iϕj (m+1) and Nm−1 = e−iϕj (m−1) . In addition, we observe that by introducing the complementary rotation ∗ χ∗j = χ ¯∗j eiϕj , a similar equation with complex conjugated coefficients, s∗ (ω 2 ), is obtained for the conjugated roots χ ¯∗j (Figure 3.10). Indeed, a straightforward algebraic manipulation yields a characteristic equation coalescing equation [3.49]. The sign of the real parts χ ¯jR of the roots of equation [3.49], corresponds henceforth to the roots χj of the characteristic equation in ¯ equation [3.42] that may be: i) Re(χ ¯j ) < 0 ∀j ⇒ arg(χj ) ≥ |βπ| ∀j; 2 ¯

ii) Re(χj ) = 0 ⇒ arg(χj ) = |βπ| 2 ∀j. Previous consideration yields that phase shift of the stability sector in the complex plane represents a conformal mapping of the C plane with one to one correspondence. The aforementioned considerations suggest that the stability of fractional-order dynamical systems may be studied via the Routh–Hurwitz criterion applied to the complex characteristic equation in equation [3.49]. The Routh–Hurwitz criterion will be applied over the real coefficient characteristic equation S 2 (ω 2 (λ)) = 0 of order 2m, obtained by the product

Analysis of a Beck’s Column over Fractional-Order Restraints

63

S 2 (ω 2 (λ)) = s(ω 2 )s∗ (ω 2 ). Signs of the real part of the roots of the equation S 2 (ω 2 (λ)) = 0 can be obtained by the sign of the sequence of determinants. p = V (1, Δ1 , . . . , Δn )

[3.50a]

q = V (1, −Δ1 , . . . , (−1)n Δn )

[3.50b]

where Δk (Marden 1949) has the structure of: m1 m3 · · · m2k−1 −n2 −n4 · · · 1 m2 · · · m2k−2 −n1 −n3 · · · 0 m1 · · · m2k−3 0 −n2 · · · .. .. . 0 1 ··· . . 0 .. Δk = 0 0 ··· 0 0 · · · m2k 0 n2 · · · n2k−2 m1 m3 · · · . .. n1 · · · n2k−3 1 m2 · · · 0 0 · · · nk 0 0 ···

−n2k−2 −n2k−3 −n2k−4 .. . −nk−1 −mk−3 .. . −mk−1

[3.51]

for k = 1, ·, n, with mj and nj , with j = 1, ·, n are the coefficients of the characteristic polynomial, mj are the coefficients of real part and nj are the coefficients of the imaginary part.

Figure 3.10. Stability region and complex map

3.5. Numerical application We report our findings in Figures 3.11–3.13 for the cases that we analyzed, and we can conclude that the presence of a fractional Kelvin–Voigt model, a Fractional

64

Modern Trends in Structural and Solid Mechanics 1

Zener, in which the spring with stiffness k1 → 0, keeping Cβ = 0, does not remove the Hermann–Smith and Ziegler paradox as already observed in the classical approach to the problem.

Figure 3.11. Plot of the critical load for β = 0.0001

Figure 3.12. Plot of the critical load for β = 0.14

Parametric analysis shows the influence of the parameter τs and the effect of the order of derivation on the stability value of the applied load. It can be observed from the reported parametric analysis that as β = 10−4 (β → 0 external elastic restraints), the load λ is almost insensitive to the presence of the external restraints, as predicted by the Hermann–Smith paradox. The analysis has been carried out introducing the β anomalous time scaling τβ = (τs ) for the abscissa and material parameter.

Analysis of a Beck’s Column over Fractional-Order Restraints

65

Figure 3.13. Plot of the critical load for β = 1

In the case of the viscous restraint, the common value λ = 4.13 is obtained as τs → 0 for any value of the differentiation order β, as predicted by the Ziegler paradox (Ziegler 1952). 3.6. Conclusion The analysis of Beck’s column resting on an external foundation is a challenging and still unsolved problem in the field of mechanics. Indeed, the presence of well-established paradoxes arising from stability analysis and involving the evaluation of critical loads yields unreliable solutions in terms of stability. In this study, we aimed to replace the external restraints with a generalized version of material behavior involving fractional-order calculus that proved to be very efficient in representing real material behavior. In the presence of fractional-order constitutive equations, the usual version of the Routh–Hurwitz criterion, based on the sign of the real part of the eigenvalues of the dynamic matrix of the problem, cannot be used since the stability region is not the entire half-positive complex plane. In this study, we also showed that in the presence of more general foundations, besides the classical fractional type, the Routh–Hurwitz theorem may be generalized by introducing a conformal mapping of the boundaries of the stability region that moves to the complex-half plane. Some numerical results assessing the reliability of the proposed approach have been shown in the study. 3.7. References Atanackovic, T.M. and Stankovic, B. (2004). On a system of differential equations with fractional derivatives arising in rod theory. Journal of Physics A: Mathematical and General, 37(4), 1241. Atanackovic, T.M., Janev, M., Konjik, S., Pilipovic, S., Zorica, D. (2015). Vibrations of an elastic rod on a viscoelastic foundation of complex fractional Kelvin–Voigt type. Meccanica, 50(7), 1679–1692.

66

Modern Trends in Structural and Solid Mechanics 1

Barnett, S. (1970). Greatest common divisor of two polynomials. Linear Algebra and its Applications, 3(1), 7–9. Beck, M. (1952). Die knicklast des einseitig eingespannten, tangential gedr¨uckten stabes. Zeitschrift f¨ur angewandte Mathematik und Physik ZAMP, 3(3), 225–228. Bologna, E. and Zingales, M. (2018). Stability analysis of Beck’s column over a fractional-order hereditary foundation. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, 474(2218), 20180315. Bologna, E., Deseri, L., Zingales, M. (2017). A state-space approach to dynamic stability of fractional-order systems: The extended Routh-Hurwitz theorem. AIMETA2017, Salerno, 4–7 September. Bologna, E., Graziano, F., Deseri, L., Zingales, M. (2019). Power-laws hereditariness of biomimetic ceramics for cranioplasty neurosurgery. International Journal of Non-Linear Mechanics, 115, 61–67. Bologna, E., Di Paola, M., Dayal, K., Deseri, L., Zingales, M. (2020). Fractional-order nonlinear hereditariness of tendons and ligaments of the human knee. Philosophical Transactions of the Royal Society A, 378(2172), 20190294. Elishakoff, I. (2001). Euler’s problem revisited: 222 years later. Meccanica, 36(3), 265–272. Elishakoff, I. (2005). Controversy associated with the so-called “follower forces”: Critical overview. Applied Mechanics Reviews, 58(2), 117–142. Mainardi, F. (2010). Fractional Calculus and Waves in Linear Viscoelasticity: An Introduction to Mathematical Models. Imperial College Press, London. Marden, M. (1949). Geometry of Polynomials. American Mathematical Society, Providence, RI. Matignon, D. (1996). Stability results for fractional differential equations with applications to control processing. Multiconference on Computational Engineering in Systems Applications, vol. 2, Lille, France, 963–968. Pinnola, F.P. (2016). Statistical correlation of fractional oscillator response by complex spectral moments and state variable expansion. Communications in Nonlinear Science and Numerical Simulation, 39, 343–359. Plaut, R.H. and Infante, E. (1970). The effect of external damping on the stability of Beck’s column. International Journal of Solids and Structures, 6(5), 491–496. Podlubny, I. (1998). Fractional Differential Equations: An Introduction to Fractional Derivatives, Fractional Differential Equations, to Methods of their Solution and Some of their Applications. Academic Press, San Diego, CA. Rossikhin, Y.A. and Shitikova, M. (2001). A new method for solving dynamic problems of fractional derivative viscoelasticity. International Journal of Engineering Science, 39(2), 149–176. Smith, T. and Herrmann, G. (1972). Stability of a beam on an elastic foundation subjected to a follower force. Journal of Applied Mechanics, 39(2), 628–629. Spanos, P.D. and Miller, S.M. (1994). Hilbert transform generalization of a classical random vibration integral. Journal of Applied Mechanics, 61, 575–581. Ziegler, H. (1952). Die stabilit¨atskriterien der elastomechanik. Ingenieur-Archiv, 20(1), 49–56.

4 Localization in the Static Response of Higher-Order Lattices with Long-Range Interactions

The static behavior of a generalized axial lattice with direct and indirect neighboring interactions is studied under uniform tensile loading. The two-neighbor interaction lattice is composed of different elastic springs connected to adjacent nodes and to next adjacent nodes, with possible different stiffness values for each interaction. The boundary nodes of this one-dimensional generalized lattice are assumed to be fixed for one end and loaded by a prescribed elongation at the free other end. The static response of this generalized axial lattice under uniform tensile load is analytically studied, from the resolution of a linear fourth-order difference equation. It is shown that the so-called higher-order boundary conditions have a crucial influence on the possible localization process of the displacement in the vicinity of the clamped end. The referenced homogeneous response may be perturbed when considering non-symmetric higher-order boundary conditions at the fixed nodes (or truncated higher-order boundary conditions). This non-symmetric higher-order boundary condition is also interpreted in terms of local weakening of the stiffness in the vicinity of the clamped section. A non-monotonic boundary layer is exhibited for the discrete displacement field of this generalized lattice. This discrete boundary layer is analyzed from an asymptotic approach, and is approximated through a simplified formulae based on the exponential law, with a change of sign from one particle to the next.

Chapter written by Noël CHALLAMEL and Vincent PICANDET. Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

68

Modern Trends in Structural and Solid Mechanics 1

4.1. Introduction Periodic subsystems, which possess periodic or almost periodic microstructures, are basic structural elements that may be introduced in a refined representation of many physical and mechanical problems (see Banakh and Kempner 2010 or Wang et al. 2020). The microstructure is controlled by some characteristic sizes with respect to the structural scale. Microstructure effects may play a decisive role in the appearance of some localization phenomena. A possible approach to describing these periodic microstructures is based on lattice mechanics, formulated in terms of discrete cells, connected by some short- or long-range interactions. The wave dispersive properties of these periodic lattices are well studied, and are available in standard textbooks devoted to the wave propagation in heterogeneous or microstructured media (Brillouin 1946; Born and Huang 1954; Aşkar 1986; Maugin 1999; Craster and Kaplunov 2013). In this chapter, we will focus on linear elastic interactions, for each particle, with restoring force based on the interaction with the adjacent particle, as well as with its second neighboring. This lattice may be labeled as a generalized lattice with a general interaction dependence between particles next to the considered adjacent particle. Such a long-range lattice is the generalization of simple lattices considered by Lagrange (1759, 1788) (one-dimensional string or axial lattices) or by Born and von Kármán (1912) (including three-dimensional elastic lattices) based on direct neighboring interaction. The wave dispersive properties in a generalized infinite lattice (with short- and long-range interactions) have been analytically studied by Brillouin (1946), Eaton and Peddieson (1973) and Roseneau (1987). As described in the seminal book by Maugin (1999), lattices with multiple short- and long-range interactions may play a key role in the understanding of non-local and nonlinear generalized wave propagation, whose stability property may be sensitive to the interaction complexity (see more recently Tarasov (2010) or Michelitsch et al. (2019)). Generalized lattices may also be strongly related to integral-based non-local elastic media (see, for instance, Eringen and Kim (1977)). The vibration properties of such generalized elastic lattices of finite size have been extensively studied during the last decades, with closed-form solutions for linear elastic interactions. Pipes (1966), Chen (1970, 1971), Eaton and Peddieson (1973) or more recently Challamel et al. (2018, 2019) calculated the natural frequencies of a finite generalized lattice (with direct and indirect neighboring interactions). Pipes (1966) and Chen (1970, 1971) considered a specific finite lattice with N-neighbor interaction characterized by equal stiffness for each interaction. Eaton and Peddieson (1973) obtained the eigenfrequencies of a fixed–fixed, fixed–free and free–free generalized lattice with N-neighbor interaction, with possible differentiation between short- and long-range elastic interactions. Eaton and Peddieson (1973) also approximated the dynamic equation of the generalized lattice (with short- and long-range interactions) by a continuous rod

Localization in the Static Response of Higher-Order Lattices

69

model with higher-order derivatives (higher-order gradient elasticity model). Carcaterra et al. (2015) related the topology of the microstructure connections of the generalized lattice (with short- and long-range interactions) to some asymptotic properties of a gradient continuous medium (with convergence properties for sufficiently small lattice spacing). Challamel et al. (2018) recently reconsidered the vibration behavior of this finite generalized lattice with relevant scaling parameters, based on an asymptotic behavior of the generalized lattice which converges towards the continuous local rod model for sufficiently small spacing parameters. Challamel et al. (2018), following the works of Rosenau (1987), also built an approximated continuous non-local rod model without additional higher-order spatial derivatives for capturing the length scale dependence of the generalized lattice. This non-local rod model is shown to be equivalent to a stress gradient Eringen model (in the sense of the differential model of Eringen (1983). The eigenfrequencies of this non-local rod model fit the exact ones of the generalized lattice with strong accuracy. The scale dependence of the eigenfrequencies of this generalized lattice is found to be controlled by the lattice spacing and the stiffness ratio of the direct and indirect neighboring interactions. Challamel et al. (2018) (see also Challamel et al. 2019) showed that the calculation of the eigenfrequency of the generalized lattice needs to solve a higher-order difference boundary value problem. The eigenfrequency solutions of Eaton and Peddieson (1973) have been confirmed again in the presence of symmetrical or antisymmetrical higher-order boundary conditions. However, for some unsymmetrical higher-order boundary conditions, the scale dependence of the eigenfrequencies may be drastically affected, thus showing the strong dependence of the mechanical problem to the choice of higher-order boundary conditions. This strong dependence of the mechanical response to the choice of higher-order boundary conditions is revisited in this chapter, and restricted to a static analysis for a generalized lattice under uniform tensile loading. Exact resolution of linear lattice problems can be attained from the mathematical properties of linear difference equations (see the monographs of Goldberg 1958 or Elaydi 2005). Exact solutions of axial lattice with direct and indirect neighboring interactions have been obtained in statics by Gazis and Wallis (1964), Toupin and Gazis (1964) and Mindlin (1965). Mindlin (1965) also included the influence of a third neighboring interaction. Toupin and Gazis (1964) considered the static behavior of a lattice chain with direct and next-to-direct elastic interaction, loaded by a system of axial tensile loads. They pointed out the existence of a boundary layer for this problem, and the possible coupling exponential regime for the displacement with a change of sign from one particle to the next. Triantafyllidis and Bardenhagen (1993) obtained numerical solutions for a nonlinear axial lattice under uniform axial load and with possible interaction with direct or other adjacent elements. Such a nonlinear discrete problem is the generalization of FPU nonlinear lattices (Fermi et al. 1955) with both direct and next-to-direct interactions. Very few analytical solutions are available for the static response of nonlinear lattices (see

70

Modern Trends in Structural and Solid Mechanics 1

Hérisson et al. (2016) for the derivation of exact analytical solutions of the static response of FPU systems under non-uniform distributed axial load). Triantafyllidis and Bardenhagen (1993) and Hérisson et al. (2016) characterized the localization phenomenon due to the nonlinear restoring force between adjacent particles (these last problems can be viewed as the discrete investigation of the nonlinear Ericksen (1975) continuous bar). This localization phenomenon is different from the one considered in this chapter where only linear elastic interactions are considered. In this chapter, it will be shown that even in the presence of pure linear interactions, the specific effect of additional long-range interaction may be responsible for some localization phenomena. Charlotte and Truskinovsky (2002, 2008) investigated the static behavior of a lattice composed of direct and next-to-direct linear elastic interaction and showed the strong influence of discrete-based boundary conditions, through a soft and a hard loading device (displacement- or force-based boundary conditions). Charlotte and Truskinovsky (2002) also performed a generic study of the mathematical property of the associated linear difference equation of the generalized lattice, with possible negative sign for the generalized stiffness. In this chapter, the static response of a generalized lattice with direct and next-to-direct elastic interactions is studied. An exact solution is developed for this generalized lattice under uniform tensile load with symmetrical and unsymmetrical higher-order boundary conditions at the clamped end. This solution is obtained from the resolution of a linear fourth-order difference equation. For unsymmetrical higher-order boundary conditions, an oscillatory boundary layer is exhibited in the vicinity of the clamped end. This boundary layer, specific from the discreteness of this lattice problem, is analytically characterized from an asymptotic approach, with oscillating terms from one node to another. 4.2. Two-neighbor interaction – general formulation – homogeneous solution The generalized axial lattice with two-neighbor interactions is considered, as shown in Figure 4.1. This lattice is composed of (n+1) concentrated masses (or particles). The axial displacement of a mass at node i is denoted by ui. The spacing between each mass is assumed to be uniform and is denoted by a. The axial spring stiffness of the direct neighboring interaction is denoted by k1 , whereas the axial spring stiffness of the second neighboring interaction is denoted by k2 . The lattice is composed of equal concentrated massed m attached at each node, except at the end nodes of the lattice. The fixed boundary conditions is ensured by fixing the first node of the lattice and by assuming that the second

Localization in the Static Response of Higher-Order Lattices

71

neighboring is attached to the fixed node with a spring stiffness 2k2 that is twice the stiffness value k2 inside the lattice. Such a stiffness calibration is similar to that considered for calibrating the rotational stiffness at the lattice clamped boundary condition of the Hencky system (see Hencky (1920) who already mentioned that the calibrated stiffness of the clamped section is twice the one inside the beam domain, or see Challamel et al. (2014) and Wang et al. (2017) for a recent discussion of this problem in conjunction with a non-local beam mechanics insight).

Figure 4.1. Extension of a generalized lattice with direct and indirect interactions under displacement controlled test un = un ; symmetrical higher-order boundary conditions at the clamped section u−1 = −u1 ; n=4; uniform loading

The balance equation can be obtained from the following difference equation:

Ni +1 2 − N i −1 2 = 0 with N i +1 2 = k1 ( ui +1 − ui ) + k 2 ( ui + 2 − ui ) + k2 ( ui +1 − ui −1 )

[4.1]

It is possible to introduce the following scaling law for the stiffness: k1 = β1

EA EA , k2 = β 2 a 4a

[4.2]

with the following parameter constraints β1 ≥ 0 , β 2 ≥ 0 . EA is the axial rigidity of the continuous axial bar, asymptotically obtained for a number n of nodes tending to infinite. We also have β1 + β 2 = 1 so that the parameters β1 and can be considered as weighting coefficients for the short- and long-range interactions. Moreover, it is

72

Modern Trends in Structural and Solid Mechanics 1

physically reasonable to assume that the weighting positive coefficients βi decrease for the long-range interaction, which means that:

β1 ≥ β 2 ≥ 0 

1 1 ≤ β1 ≤ 1 and 0 ≤ β 2 ≤ 2 2

[4.3]

The difference equations can be reformulated as:

Ni +1 2 − Ni +1 2 a

=0

with

u −u u −u   u −u Ni +1 2 = EA  β1 i +1 i + β 2 i + 2 i + β 2 i +1 i −1  4a 4a  a 

[4.4]

The mixed differential–difference equation is obtained by coupling both equations:

u − 2ui + ui − 2   u − 2ui + ui −1 EA  β1 i +1 + β2 i +2 2 =0 a 4a 2  

[4.5]

In this chapter, the case β 2 ≥ 0 will be investigated, which is physically motivated by the positivity of the associated lattice potential energy. Gazis and Wallis (1964) or Toupin and Gazis (1964) remark that in this case, the amplitude of the displacement varies exponentially with the position of the particle, but the sign changes from one particle to the next (in a boundary layer). Gazis and Wallis (1964) and Toupin and Gazis (1964) also considered the case β 2 ≤ 0 with exponential decaying in the boundary zone (but this case is associated with a non-definite positive energy function of the generalized lattice). This difference equation (equation [4.5]) may have been obtained equivalently from energy arguments. For instance, starting from the fixed–fixed lattice system, the potential energy of the linear elastic lattice with two closest neighbors may be written as: n −1

W = i =0

n−2 k1 k 2 2 2 2 ( ui +1 − ui ) +  2 ( ui + 2 − ui ) + k2 ( u1 − u0 ) + k2 ( un − un −1 ) 2 2 i =0

[4.6]

Localization in the Static Response of Higher-Order Lattices

73

with the following parameter constraints k1 ≥ 0 , k2 ≥ 0 so as to ensure the definite positivity of the potential energy. For a fixed boundary condition at the clamped section, and prescribed displacement at the other end un = un (uniaxial tensile loading test), the kinematic constraints are imposed such that: u0 = 0 and un = un

[4.7]

Equation [4.6] can also be equivalently written in terms of fictitious nodes outside the domain: n −1

W = i =0

n−2 k1 k k k 2 2 2 2 ( ui +1 − ui ) +  2 ( ui + 2 − ui ) + 2 ( u1 − u−1 ) + 2 ( un +1 − un −1 ) [4.8] 2 2 4 4 i =0

with the antisymmetrical higher-order additional boundary conditions: u−1 = −u1 and un +1 = 2un − un −1

[4.9]

The potential energy of the linear elastic lattice with two closest neighbors is then written as: n −1

W =  β1 i =0

2

2

n−2 EA  ui +1 − ui  EA  ui + 2 − ui  × a +  β2 ×a   2  a  2  2a  i =0 2

2

u −u   u − un −1  + β 2 EA  1 0  × a + β 2 EA  n  ×a  2a   2a 

[4.10]

For the prescribed displacement-controlled test, the tensile load is calculated from the discrete displacement field solution:

(

)

(

)

(

F = k1 un − un −1 + 2k2 un − un −1 + k2 un − un − 2

)

[4.11]

The dimensionless parameters of the problem can be introduced as: F* =

u F , ui* = i L EA

[4.12]

so that the dimensionless load–displacement relationship at the end of the generalized lattice can be written as:

(

)

(

β  β  F * = n  β1 + 2  un* − un*−1 + n 2 un* − un*− 2 2 4  

)

[4.13]

74

Modern Trends in Structural and Solid Mechanics 1

This linear fourth-order difference equation [4.5] can be equivalently rewritten as: ui +1 − 2ui + ui −1 +

β2 4

( ui + 2 − 4ui +1 + 6ui − 4ui −1 + ui − 2 ) = 0

[4.14]

which can be presented in a condensed formula:  a2 β2  Δ 1 + Δ  ui = 0 4  

[4.15]

where Δ is the discrete Laplacian. Equation [4.15] directly shows that the solution of this fourth-order difference equation can be decomposed into a long wave and a boundary layer contribution: ui = ui + uˆi

[4.16]

where the long-wave solution is governed by a discrete Laplace equation: Δ ui = 0

[4.17]

whose solution is linear ui = A1 + A2 i

[4.18]

whereas the second solution is obtained from a second-order difference equation:  a2 β2  Δ  uˆi = 0 1 + 4  

[4.19]

The solution of this linear second-order difference equation is sought in the form (see Goldberg 1958; Elaydi 2005)

uˆi =



p∈{1,2}

C p f pi

[4.20]

where the two solutions of the characteristic equation f p are detailed below  4  − 2 f +1 = 0 f 2 +  β2 

[4.21]

Localization in the Static Response of Higher-Order Lattices

75

whose solutions may be expressed by: f1,2 = 1 −

2

 2  ± 1 −  − 1 β2  β2  2

[4.22]

Recall that β 2 ∈ [ 0;1] (and in fact for physical reasons, it is reasonable to assume that β 2 ≤ β1 ), so that the solutions may be equivalently reexpressed by:  β  f1,2 = − cosh θ ± sinh θ with θ = a cosh  1 + 2 1  β2  

[4.23]

The solution to equation [4.14] can be finally written as the sum of the two solutions:

ui = A1 + A2 i + A3 ( −1) cosh (θ i ) + A' 4 ( −1) sinh (θ i ) i

i

[4.24]

This is in accordance with the solution derived by Challamel et al. (2019) for dynamic problems, when the inertia effects are neglected. This is also the solution obtained by Toupin and Gazis (1964) with symmetry considerations A1 = A3 = 0 . Toupin and Gazis (1964) also investigated the case β 2 ≤ 0 with a purely exponential long-range solution, but without physical justification of such a sign (in this chapter, we only treat the case β 2 ≥ 0 ). Charlotte and Truskinovsky (2002) also presented a parametric study of the generic solutions of the fourth-order difference equation [4.5], and also obtained a similar solution expressed by equation [4.24] for positive stiffness parameters (which is the physically based constrained parameters of the problem). The four boundary conditions associated with the case presented in Figure 4.1 are: u0 = 0 , u−1 = − u1 , un = un and un +1 − un = un − un −1

[4.25]

Inserting the solution presented in equation [4.24] in the four boundary conditions (equation [4.25]) gives the four constants of the solution: A1 = A3 = A4 = 0 and A2 =

un n

[4.26]

76

Modern Trends in Structural and Solid Mechanics 1

which means that the discrete displacement field is given for this problem as: ui =

i un n

[4.27]

The unperturbed homogeneous solution is obtained in this case. Using equation [4.27], the load–displacement relationship is obtained F=

k1 + 4k 2 EA un = un n L



F * = un*

[4.28]

This solution can also be developed from the difference equations: i = 0 : u0 = 0 i = 1: β1 ( u2 − 2u1 ) +

β2 4

( u3 − 3u1 ) = 0

i ∈ {2,..., n − 2} : β1 ( ui +1 − 2ui + ui −1 ) + i = n − 1: β1 ( un − 2un −1 + un − 2 ) + i = n : un = un

β2 4

β2 4

( ui + 2 − 2ui + ui − 2 ) = 0

( 2un − 3un −1 + un −3 ) = 0 [4.29]

4.3. Two-neighbor interaction – localization in a weakened problem

For symmetrical higher-order boundary conditions u−1 = −u1 , the spring associated with the indirect interaction in the vicinity of the fixed point has a stiffness 2k2. For the displacement controlled test, this configuration leads to the homogeneous solution with a linearly varying discrete displacement field. u−1 = 0 for the so-called truncated (or unsymmetrical) higher-order boundary conditions (see Figure 4.2). In this case, the spring associated with the indirect interaction in the vicinity of the fixed point has a stiffness k2 (instead of 2k2 considered in the previous case). This case can also be viewed as a perturbation of the homogeneous one with a local weakening in the vicinity of the clamped section.

Localization in the Static Response of Higher-Order Lattices

77

The difference equation [4.5] is still valid and may have been obtained equivalently from energy arguments, with a similar potential energy of the linear elastic lattice with two closest neighbors corrected at the boundaries at: n −1

W = i =0

n−2 k1 k k 2 2 2 2 ( ui +1 − ui ) +  2 ( ui + 2 − ui ) + 2 ( u1 − u0 ) + k2 ( un − un −1 ) [4.30] 2 2 i =0 2

with the prescribed displacements at the end nodes given by equation [4.7].

Figure 4.2. Extension of a generalized lattice with direct and indirect interactions under displacement controlled test un = un ; truncated higher-order boundary conditions at the clamped section; n=4; u−1 = 0 ; uniform loading

Equation [4.30] can also be equivalently written in terms of fictitious nodes outside the domain: n −1

W = i =0

n−2 k1 k k k 2 2 2 2 ( ui +1 − ui ) +  2 ( ui + 2 − ui ) + 2 ( u1 − u−1 ) + 2 ( un +1 − un −1 ) 2 2 4 i =0 2

[4.31]

with the truncated (or unsymmetrical) higher-order additional boundary conditions: u−1 = 0 and un +1 = 2un − un −1

The solution can also be presented by the following difference equations: i = 0 : u0 = 0

[4.32]

78

Modern Trends in Structural and Solid Mechanics 1

i = 1: β1 ( u2 − 2u1 ) +

β2 4

( u3 − 2u1 ) = 0

i ∈ {2,..., n − 2} : β1 ( ui +1 − 2ui + ui −1 ) + i = n − 1: β1 ( un − 2un −1 + un − 2 ) +

β2 4

β2 4

( ui + 2 − 2ui + ui − 2 ) = 0

( 2un − 3un −1 + un −3 ) = 0

i = n : un = un

[4.33]

The fourth-order difference equation [4.14] is still valid (only the higher-order boundary conditions differ between the two cases represented in Figures 4.1 and 4.2). The solution is expressed by equation [4.24], whose four constants can be identified from the four boundary conditions given by equation [4.34] for the problem with an unsymmetrical higher-order boundary condition at the clamped section (which can be viewed as a perturbed case of the homogeneous one): u0 = 0 , u−1 = 0 , un = un and un +1 − un = un − un −1

[4.34]

The four constants ( A1 , A2 , A3 , A4 ) can be calculated as:             

A1 un A2 un A3 un A4 un

= = = =

sinh (θ n )

sinh (θ n ) + n sinh (θ n ) + sinh (θ n + θ )  sinh (θ n ) + sinh (θ n + θ )

sinh (θ n ) + n sinh (θ n ) + sinh (θ n + θ )  − sinh (θ n )

[4.35]

sinh (θ n ) + n sinh (θ n ) + sinh (θ n + θ )  cosh (θ n )

sinh (θ n ) + n sinh (θ n ) + sinh (θ n + θ ) 

We have the remarkable property: A3 = − A1 and A3 cosh (θ n ) + A4 sinh (θ n ) = 0

[4.36]

Localization in the Static Response of Higher-Order Lattices

79

The load–displacement relationship can be calculated in this case from equation [4.13], which may highlight the reduced stiffness of this weakened system: F* u

* n

=n

A2 un

= 1+

1 sinh (θ n )

≤1

[4.37]

n sinh (θ n ) + sinh (θ n + θ ) 

The reduced stiffness in equation [4.37] can be approximated with a strong accuracy using the asymptotic property: lim

n →∞

sinh (θ n )

sinh (θ n ) + sinh (θ n + θ )

=

1 e −θ = 1 + eθ 1 + e −θ

[4.38]

Furthermore, the remarkable property can be used:

1 − β1 e−θ 1 1 = = = −θ θ 2 1+ e 1+ e 2 1 + cosh θ + cosh θ − 1 with cosh θ =

1 + β1 1 − β1

[4.39]

Figure 4.3. Displacement field in the generalized lattice under uniform loading – n=4 – uniform loading – truncated higher-order boundary conditions; 1=0.01; 2=0.99

80

Modern Trends in Structural and Solid Mechanics 1

Figure 4.4. Displacement field in the generalized lattice under uniform loading – n=10 – uniform loading – truncated higher-order boundary conditions; 1=0.01; 2=0.99

It can be concluded that the reduced stiffness can be evaluated with a simplified formulae: lim

n →∞

FL = EAun

1 1 − β1 1+ 2n

≤1

[4.40]

Based on 1=0.01; 2=0.99 (asymptotic case valid for a lattice with predominant indirect interactions), Figure 4.3 shows, for a small number of elements n=4, the displacement field in the lattice with the so-called truncated higher-order boundary conditions (or unsymmetrical higher-order boundary conditions), when compared to symmetrical higher-order boundary conditions. The inhomogeneous displacement field is clearly shown with such higher-order boundary conditions, due to the second part of the solution in exponential form (see also Figure 4.4 with a more pronounced oscillatory behavior in the vicinity of the clamped section).

Localization in the Static Response of Higher-Order Lattices

81

Figure 4.5. Displacement field in the generalized lattice under uniform loading – n=4 – uniform loading – truncated higher-order boundary conditions; influence of the dimensionless coefficient β1 ∈ {0.1;0.5;0.9}

Note that the values 1=0.01 and 2=0.99 are quite extremal, because such higher-order lattice is mainly governed by next-to-direct interactions instead of direct interactions. The localization phenomenon is characterized by some oscillation terms in the vicinity of the clamped section (or reduced section). Gazis and Wallis (1964) and Toupin and Gazis (1964) already remarked that in a similar problem, the amplitude of the displacement varies exponentially with the position of the particle, but the sign changes from one particle to the next (in a boundary layer). It is worth mentioning that Charlotte and Truskinovsky (2002) also concluded that for such a physically based problem (with positive short- and long-range stiffness parameters) that the boundary layer contains some oscillations at the scale of the lattice modulated by an exponential envelope. The localization phenomenon in the vicinity of the fixed point is even more pronounced for a larger number of elements where an oscillatory behavior is observed (see Figure 4.4). In Figure 4.4, n has been chosen equal to n=10, when compared to n=4 in Figure 4.4 with the same weighting coefficients 1=0.01 and 2=0.99.

82

Modern Trends in Structural and Solid Mechanics 1

Figure 4.6. Equivalent stiffness versus n; uniform loading – truncated higher-order boundary conditions; β1 = 0.5

Figure 4.5 shows that the localization phenomenon is more pronounced for predominant next-to-direct interactions (small values of β1 and large values of β 2 with 0 ≤ β1 ≤ β 2 ≤ 1 ). When β1 approaches unity, the solution in the case of unsymmetrical boundary conditions asymptotically converges towards the homogeneous solution valid for the symmetrical higher-order boundary condition. The inhomogeneous solution tends towards the homogeneous one for a sufficiently large number n of elements, as shown in Figure 4.6. In this part, an approximated solution will be investigated for a better understanding of the localization phenomenon controlled by the boundary layer in the vicinity of the fixed point. We here employed a methodology followed by Kaplunov and Pichugin (2009) for evaluating the nature of the boundary layer in the vicinity of the clamped section for the so-called perturbed case (with truncated higher-order boundary conditions). Kaplunov and Pichugin (2009) calculated a boundary layer solution for the dynamic response of a lattice model with direct neighboring interactions and displacement-based boundary conditions (as considered in the present study restricted to a static analysis). It is worth mentioning that this boundary layer is specific to the discreteness of the lattice, characterized by a strong oscillatory regime controlled by the size of the lattice spacing.

Localization in the Static Response of Higher-Order Lattices

83

The exact displacement field is decomposed into a long wave and a boundary layer contribution: ui = ui + uˆi with ui = A1 + A2 i

[4.41]

The boundary layer solution is reexpressed for convenience in the following way: uˆi = ( −1) C1eθ i −θ n −θ + C2 e −θ i −θ  i

[4.42]

Now, the two boundary conditions for the unsymmetrical higher-order boundary condition at the clamped section are written with respect to these two displacement fields: uˆ ( 0 ) + u ( 0 ) = 0 and uˆ ( − a ) + u ( − a ) = 0

[4.43]

which can also be reexpressed as: u ( 0 ) = −uˆ ( 0 ) and u ( 0 ) − au ′ ( 0 ) = −uˆ ( − a )

[4.44]

The use of the boundary layer solution [4.42] in equation [4.44] leads to the two asymptotic boundary conditions: u ( 0 ) = − C1e−θ n −θ + C2 e−θ  ~ −C2 e−θ

and u ( 0 ) − au ′ ( 0 ) = C1e −θ n − 2θ + C2  ~ C2

[4.45]

which means that the long-wave boundary condition is written in a single equation as: u ( 0 ) + u ( 0 ) − au ′ ( 0 )  e−θ = 0

[4.46]

The boundary condition at the end associated with the application of the prescribed displacement (and implicitly the boundary layer is assumed to be negligible in this zone):

u ( L ) = un

[4.47]

84

Modern Trends in Structural and Solid Mechanics 1

Using both equations [4.46] and [4.47] gives the constant of the so-called long-wave solution (approximated solution): un

A1 =

L+a

−θ

e 1 + e −θ

ae −θ and A2 = 1 + e −θ

un L+a

e −θ 1 + e −θ

[4.48]

Using the mathematical property equation [4.39], the long-wave solution (slow varying solution) is then written as: 1 − β1 i + n ui = un 2n 1 − β1 +1 2n

[4.49]

For similar reasons, the boundary layer solution can be restricted asymptotically to the second decaying exponential term in equation [4.42]: uˆi ~ ( −1)

i +1

u ( 0 ) e −θ i

[4.50]

so that an approximated solution may be obtained for the case with unsymmetrical higher-order boundary conditions: 1 − β1 1 − ( −1)i e−θ i  + i   n n 2 = un 1 − β1 +1 2n ui

[4.51]

Figure 4.7 shows that this approximate solution efficiently fits the exact lattice solution, especially in the vicinity of the fixed point. It is confirmed from equation [4.51] that the boundary layer term is negligible in the case of pure direct interactions (1=1), whereas the contribution of the boundary layer is predominant for small values of 1 (and large values of 2 which controls the next-to-direct interaction contribution). It is worth mentioning that an alternative boundary condition at the prescribed end would be based on the expansion of equation [4.11]:  β  u − un −1 β 2 un − un − 2  + F = EA  β1 + 2  n  a 2  2 2a  

[4.52]

Localization in the Static Response of Higher-Order Lattices

85

which can be expanded leading to the first-order approximation: u ′ ( L ) =

F EA

[4.53]

Figure 4.7. Displacement field in the generalized lattice under uniform loading – n=10 – uniform loading – truncated higher-order boundary conditions; 1=0.01; 2=0.99; approximated boundary layer solution

In this case, we obtain: ui =

FL  1 − β1 i  +   EA  2n n 

[4.54]

The boundary layer contribution can be added in the total displacement formulation, leading to the same expression [4.51]. The axial stiffness is reduced,  1 − β1  according to equation [4.54] from a factor 1  1 +  , which is also consistent  2n   with the asymptotic expansion used to derive equation [4.40]. This result is also not in contradiction with the weakening found by Challamel et al. (2019) for the eigenfrequency of a fixed–fixed lattice with direct and next-to-direct interaction. Based on a truncated higher-order boundary condition at both sides, Challamel et al. (2019) found from an asymptotic expansion a weakening

86

Modern Trends in Structural and Solid Mechanics 1

of the fundamental frequency from a factor 1 −

1 − β1

. It is seen that the reduced n stiffness factor asymptotically obtained by Challamel et al. (2019) in a dynamic analysis (for a two-side weakening problem) is two times the one obtained in the present static study (for a one-side weakening problem). 4.4. Conclusion

In this chapter, the static behavior of a generalized lattice composed of direct and indirect elastic interactions is studied under uniform tensile load. It is shown that the choice of higher-order boundary conditions strongly affects the macroscopic response of this microstructured system. The homogeneous response is recovered in the case of symmetrical higher-order boundary conditions. Unsymmetrical higher-order boundary conditions at the clamped section are responsible for some localization phenomena in the vicinity of the clamped section. A non-monotonic boundary layer is exhibited for the discrete displacement field of this generalized lattice. This discrete boundary layer is analyzed from an asymptotic approach, and is approximated through a simplified formulae based on the exponential law, with a change of sign from one particle to the next. Boundary layers have already been observed in continuous non-local media where some localization phenomena may be predominant (see, for instance, Chebakov et al. 2016). Seemingly, the discrete treatment of the boundary layers presented in this chapter is more consistent than a continuous treatment since the homogenization of the boundary layers does not appear to be fully justified, due to their short-wavelength nature. The present results could be generalized to a nonlinear higher-order lattice, as investigated numerically by Triantafyllidis and Bardenhagen (1993). This is the generalization of FPU nonlinear lattices (Fermi et al. 1955) with both direct and next-to-direct interactions. Very few analytical solutions are available for the static response of nonlinear lattices (see Hérisson et al. (2016) for the derivation of exact analytical solutions of FPU systems under non-uniform distributed axial load). It is expected that the boundary layer may also be present for nonlinear higher-order lattices with strong localization phenomena controlled by the spacing of the lattice and the structure of the interaction model. 4.5. References Aşkar, A. (1986). Lattice Dynamical Foundations of Continuum Theories: Elasticity, Piezoelectricity, Viscoelasticity, Plasticity. WorldScientific, Singapore.

Localization in the Static Response of Higher-Order Lattices

87

Banakh, L.Y. and Kempner, M.L. (2010). Vibrations of Mechanical Systems with Regular Structure. Springer, New York. Born, M. and Huang, K. (1954). Dynamical Theory of Crystal Lattices. Oxford University Press, Oxford. Born, M. and von Kármán, T. (1912). On fluctuations in spatial grids. Physikalishe Zeitschrift, 13, 297–309. Brillouin, L. (1946). Wave Propagation in Periodic Structures: Electric Filters and Crystal Lattices. McGraw-Hill, New York. Carcaterra, A., dell’Isola, F., Esposito, R., Pulvirenti, M. (2015). Macroscopic description of microscopically strongly inhomogenous systems: A mathematical basis for the synthesis of higher gradient metamaterials. Arch. Rational Mech. Anal., 218, 1239–1262. Challamel, N., Wang, C.M., Elishakoff I. (2014). Discrete systems behave as nonlocal structural elements: Bending, buckling and vibration analysis. Eur. J. Mech. A/Solids, 44, 125–135. Challamel, N., Wang, C.M., Zhang, H., Kitipornchai, S. (2018). Exact and nonlocal solutions for vibration of axial lattices with direct and indirect neighbouring interactions. J. Eng. Mech., 144(5), 04018025, 1–9. Challamel, N., Zhang, H., Wang, C.M., Kaplunov J. (2019). Scale effect and higher-order boundary conditions for generalized lattices, with direct and indirect interactions. Mech. Res. Commun., 97, 1–7. Charlotte, M. and Truskinovsky L. (2002). Linear elastic chain with a hyper-pre-stress. J. Mech. Phys. Solids, 50, 217–251. Charlotte, M. and Truskinovsky, L. (2008). Towards multiscale continuum elasticity theory. Cont. Mech. Thermodynamics, 20, 131–161. Chebakov, R., Kaplunov, J., Rogerson, G.A. (2016). Refined boundary conditions on the free surface of an elastic half-space taking into account non-local effects. Proc. R. Soc. A, 472, 20150800. Chen, F.Y. (1970). Similarity transformation and the eigenvalue problem of certain far-coupled systems. Am. J. Phys., 38(8), 1036–1039. Chen, F.Y. (1971). On modeling and direct solution of certain free vibration systems. J. Sound Vibration, 14(1), 57–79. Craster, R.V. and Kaplunov, J. (2013). Dynamic Localization Phenomena in Elasticity, Acoustics and Electromagnetism. Springer, Vienna. Eaton, H.C. and Peddieson Jr., J. (1973). On continuum description of one-dimensional lattice mechanics. J. Tenn. Acad. Sci., 18(3), 96–100. Elaydi, S. (2005). An Introduction to Difference Equations, 3rd edition. Springer, New York. Ericksen, J.L. (1975). Equilibrium of bars. J. Elast., 5, 191–201.

88

Modern Trends in Structural and Solid Mechanics 1

Eringen, A.C. (1983). On differential equations of nonlocal elasticity and solutions of screw dislocation and surface waves. J. Appl. Phys., 54, 4703–4710. Eringen, A.C. and Kim, B.S. (1977). Relation between non-local elasticity and lattice dynamics. Cryst. Latt. Def., 7, 51–57. Fermi, E., Pasta, J., Ulam, S. (1955). Studies of nonlinear problems. Report, LA-1940. Los Alamos Scientific Lab, New Mexico Gazis, D.C. and Wallis, R.F. (1964). Surface tension and surface modes in semi-infinite lattices. Surface Science, 3, 19–32. Goldberg, S. (1958). Introduction to Difference Equations with Illustrative Examples from Economics, Psychology and Sociology. Dover Publications, New York. Hencky, H. (1920). Über die angenäherte lösung von stabilitätsproblemen im raummittels der elastischen gelenkkette. Der Eisenbau, 11, 437–452. Hérisson, B., Challamel, N., Picandet, V., Perrot, A. (2016). Nonlocal continuum analysis of a nonlinear uniaxial elastic lattice system under non-uniform axial load. Physica E, 83, 378–388. Kaplunov, J.D. and Pichugin, A.V. (2009). On rational boundary conditions for higher-order long-wave models. In IUTAM Symposium on Scaling in Solid Mechanics, Borodich, F.M. (ed.). IUTAM Book, 10, 81–90. Lagrange, J.L. (1759). Recherches sur la nature et la propagation du son. Miscellanea Philosophico-Mathematica Societatis Privatae Taurinensis I, 2rd Pagination, i-112, (see also Œuvres, Volume 1, 39–148). Lagrange, J.L. (1853). Mécanique analytique, 3rd edition. Paris, 1788, Mallet-Bachelier, Gendre et successeur de bachelier, Imprimeur-libraire du bureau des longitudes, de l’école polytechnique, de l’école centrale des arts et manufactures, Paris, 367. Maugin, G.A. (1999). Nonlinear Waves in Elastic Crystals. Oxford University Press, Oxford. Michelitsch, T., Riascos, A.P., Collet, B., Nowakowski, A., Nicolleau F. (2019). Fractional Dynamics on Networks and Lattices. ISTE Ltd, London, and John Wiley & Sons, New York. Mindlin, R.D. (1965). Second gradient of strain and surface-tension in linear elasticity. Int. J. Solids Structures, 1, 417–438. Pipes, L.A. (1966). Circulant matrices and the theory of symmetrical components. Matrix Tensor Quart., 17, 35–50. Rosenau, P. (1987). Dynamics of dense lattices. Phys. Rev. B, 36(11), 5868–5876. Tarasov, V.E. (2010). Fractional Dynamics – Applications of Fractional Calculus to Dynamics of Particles, Fields and Media. Springer, New York. Toupin, R.A. and Gazis, D.C. (1964). Surface effects and initial stress in continuum and lattice models of elastic crystals. In Proc. Int. Conf. Lattice Dynamics, Wallis, R.F. (ed.). Pergamon Press, Copenhagen.

Localization in the Static Response of Higher-Order Lattices

89

Triantafyllidis, N. and Bardenhagen S. (1993). On higher order gradient continuum theories in 1-D nonlinear elasticity. Derivation from and comparison to the corresponding discrete models. J. Elast., 33, 259–293. Wang, C.M., Zhang, H., Challamel, N., Duan, W.H. (2017). On boundary conditions for buckling and vibration of nonlocal beams. Eur. J. Mech. A/Solids, 61, 73–81. Wang, C.M., Zhang, H., Challamel, N., Pan, W. (2020). Hencky-Bar-Chain/Net for Structural Analysis. World Scientific, London.

5 New Analytic Solutions for Elastic Buckling of Isotropic Plates

5.1. Introduction In this chapter, we derive a new analytical solution for determining the elastic buckling loads of thin isotropic rectangular plates with different combinations of boundary conditions. Until very recently, an analytical solution was only available for a limited number of plates with specific combinations of boundary conditions. The solution is obtained in series form, and the coefficients are solved to match the edge conditions. Thin plates are structural elements used in various fields, including civil, aeronautical and mechanical engineering. Such plates are often in a state in which distributed loading is applied normal to their edges and parallel to their planes, for example, in civil engineering, a floor slab with an applied lateral force due to wind, or, in aeronautical engineering, the panels of the lower skin of aircraft wings and fuselage under compression forces. These in-plane loads can lead to buckling, i.e. the loss of the plate’s stability. The first solution was obtained by Navier (1819), and his solution was for a simply supported plate along all four edges. For the cases where two opposite edges are simply supported, Levy (1899) developed an exact solution. For many years, these were the only cases that had an exact analytic solution for the buckling load and mode. Many methods have been used for the approximate solution of buckling in plates. These methods include the finite difference method, the finite elements method, the

Chapter written by Joseph T ENENBAUM, Aharon D EUTSCH and Moshe E ISENBERGER. Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

92

Modern Trends in Structural and Solid Mechanics 1

Rayleigh–Ritz method (Dickinson 1978; Narita and Leissa 1990), the multi-term extended Kantorovich method (Shufrin and Eisenberger 2005; Shufrin et al. 2008a, b), the differential quadrature method and a few more. In this review, we present only exact solutions. Exact analytical solutions for the elastic buckling of isotropic plates for Navier and Levy combinations of boundary conditions have been presented in several books (Timoshenko and Gere 1961; Bulson 1970; Reddy 1997; Szilard 2004; Wang et al. 2005). Recently, analytic solutions for the buckling of rectangular plates have been presented for a very limited number of new cases. Wang et al. (2016) derived a symplectic superposition method for the solution of four cases of boundary conditions (CCCC, CCCS, CCSC and CCSS). The solution is a relatively complex procedure and limited to a small number of cases. Ullah et al. (2019a) used a double finite sine integral transform method to arrive at a solution for the same four cases. In another work, Ullah et al. (2019b) derived a straightforward generalized integral transform method for finding the buckling loads of plates. In this method, they used the beam functions and arrived at a solution for the same four cases. Li et al. (2018) solved the case of a completely free plate (FFFF) with biaxial loading by the symplectic superposition method. Liu et al. (2020) used an analytical spectral flexibility method to solve three cases (CCCC, SSSS and GGGG). The complete list of the 55 possible cases for a square plate is presented in Table 5.1. This was first given by Bert and Malik (1994) for vibrations of plates. The cases include the four classical plate edge boundary conditions, which are: C – clamped, S – simply supported, F – free and G – guided. In the buckling analysis, the same 55 cases are found for unidirectional loading in the x or y direction, making a total of 110 cases. Of the 110 cases, only 100 are unique, since there are 10 cases where changing the load direction is equal to one of the cases in the other direction, and these are marked with # in Table 5.1. For cases 1–6 and 22–25, the Levy solution is known. The new solution in this chapter is only applied to cases in which all the four corners of the plate are supported, in particular, cases 7–15, 26–27, 34–35, 43–44, and 47–48. In total, we add the analytic solution for 17 cases. The cases that are not solved using the current analytic solution are marked with an * in Table 5.1. In this chapter, we present a new method of finding the exact buckling load for isotropic thin plates by using a superposition method. The solution for the buckling load of the plates is found using a static analysis. The plate is subjected to loads parallel to its plane, both uniaxial and biaxial. The solution for the buckling load of the plate is equivalent to finding the force that will yield infinite deflection at a random point of the plate. This results from the fact that buckling of a plate causes loss of stiffness and the deflection becomes infinite. A transform method from hyperbolic sine and cosine into a combination of basic trigonometric functions was used. Such

New Analytic Solutions for Elastic Buckling of Isotropic Plates

93

transformation allows us to obtain the appropriate equations needed to solve the partial differential equation of a plate. These equations satisfy the equation of equilibrium of a plate, for most of its boundary conditions. 1–6 1# SSSS 2 SSSC 3 SSSF 4 SCSC 5 SCSF 6 SFSF

7–21 7# CCCC 8 CCCS 9 CCCF 10 CSCF 11 CFCF 12 CFSF 13# CCSS 14 CCSF 15 CSSF 16*# SSFF 17* CSFF 18*# CCFF 19* SFFF 20* CFFF 21*# FFFF

22–25 22 SSSG 23 SCSG 24 SGSF 25 SGSG

26–33 26 CSCG 27 CSSG 28*# SSGG 29* CSGG 30* SSGF 31* SCGF 32* SGGF 33* SFGF

34–42 34 CGSG 35 CGCG 36* SGGG 37* CGGG 38* SGFG 39* CGFG 40* GGFG 41* FGFG 42*# GGGG

43–55 43 CCCG 44 CCSG 45* CCGF 46*# CCGG 47 CGCF 48 CGSF 49* CSGF 50* CGGF 51* CFGF 52* SGFF 53* CGFF 54*# GGFF 55* GFFF

Table 5.1. Possible combinations for boundary conditions

The analytic solution can predict the “exact” values of the critical buckling loads. But even the “exact” solutions become approximate because of the truncation of the infinite series solutions for the calculation of the numerical results. The results in this paper were obtained using 35 terms in the series, and they are converged for the number of digits that are shown. These solutions help one to understand the physical behavior at the point of buckling, and also serve as benchmark values for numerical and other approximate methods. 5.2. Equilibrium equation The equilibrium equation of in-plane stressed thin rectangular isotropic plates is (Reddy 1997)  4  ∂ w0 ∂ 2 w0 ∂ 2 w0 ∂ 4 w0 ∂ 4 w0 D −N + 2 + −N = q(x, y) [5.1] xx yy ∂x4 ∂x2 ∂y 2 ∂y 4 ∂x2 ∂y 2 where w0 (x, y) are the out-of-plane deflections of the plate, D = Eh3 /12(1 − ν 2 ) is the flexural stiffness of the plate, E is the modulus of elasticity, h is the plate thickness, and Nxx and Nyy are the pressure loads on the edges in direction x and y, respectively.

94

Modern Trends in Structural and Solid Mechanics 1

For the solution, we will also need the boundary condition along the edges of the plate. In thin plate theory, these are the deflection, slope, bending moment and shear force, along the edges parallel to the y and x axes, respectively, as (Reddy 1997) are Φx =

∂w(x, y) ∂x

[5.2]

Φy =

∂w(x, y) ∂y

[5.3]

 Mx = −D  My = −D  Qx = −  Qy = −

∂ 2 w(x, y) ∂ 2 w(x, y) + ν ∂x2 ∂y 2 ∂ 2 w(x, y) ∂ 2 w(x, y) + ν ∂y 2 ∂x2

 [5.4] 

∂w(x, y) ∂ 3 w(x, y) ∂ 3 w(x, y) D + D(2 − ν) − Nxx 3 2 ∂x ∂x∂y ∂x ∂ 3 w(x, y) ∂w(x, y) ∂ 3 w(x, y) D + D(2 − ν) − Nyy 3 2 ∂y ∂x ∂y ∂y

[5.5]  [5.6]  [5.7]

These values are calculated along the edges where either the x or y coordinate is constant. 5.3. Solution The solution for this partial differential equation is assumed as given below, and it has to satisfy the equilibrium equation all over the plate: w(x, y) =

∞  m=1

¯ m · Ym + X

∞ 

Xn · Y¯n + wp

[5.8]

n=1

The last term, wp , is the solution for a simply supported plate (SSSS) with a concentrated point load P acting at coordinates (ξ, η). The solution is given in Timoshenko (1959, p. 141). The coordinates ξ, η of the application of the load P must be chosen carefully: we try to avoid choosing a point that will be on one of the nodal lines of any of the buckling modes, as, for such a location, the vertical deflection will be always zero. One such location would be, for example, at the center of the plate, where, for some combinations of boundary condition, we will have

New Analytic Solutions for Elastic Buckling of Isotropic Plates

95

antisymmetric modes. Thus, we choose an arbitrary location for the point of application of P . The expression for wp is wp =

∞ ∞  nπη  4 P sin( mπξ a ) sin( b ) ¯ Xm · Y¯n a b dmn m=1 n=1

[5.9]

with dmn =

π2 (D π 2 b4 m4 + 2 D π 2 a2 b2 m2 n2 + D π 2 a4 n4 a4 b 4 + Nxx a4 b2 β n2 + Nxx a2 b4 m2 )

¯ m and Y¯n are trigonometric functions The two functions X     ¯ m = sin(λa x) · cos mπ + cos(λa x) · sin mπ X 2 2  nπ   nπ  + cos(λb y) · sin Y¯n = sin(λb y) · cos 2 2

[5.10]

[5.11]

[5.12]

with λa =

mπ a

[5.13]

λb =

nπ b

[5.14]

The two functions Ym and Xn are hyperbolic functions Ym = Am cosh (α1 y) + Bm cosh (α2 y) + Cm sinh (α1 y)+Dm sinh (α2 y) [5.15] Xn = En cosh (α3 x) + Fn cosh (α4 x) + Gn sinh (α3 x) + Hn sinh (α4 x) [5.16] with √  2 D(2Dπ 2 m2 + Nxx a2 β − δ1 ) α1 = 2Da √  2 D(2Dπ 2 m2 + Nxx a2 β + δ1 ) α2 = 2Da

[5.17]

[5.18]

96

Modern Trends in Structural and Solid Mechanics 1

α3 =

α4 =

√  2 D(2Dπ 2 n2 + Nxx b2 − δ2 ) 2Db

[5.19]

√  2 D(2Dπ 2 n2 + Nxx b2 + δ2 ) 2Db

[5.20]

where Nyy Nxx

[5.21]

δ1 =

 (4 π 2 m2 (β − 1) D + β 2 Nxx a2 ) a2 Nxx

[5.22]

δ2 =

 (−4 π 2 n2 (β − 1) D + Nxx b2 ) b2 Nxx

[5.23]

β=

5.4. Boundary condition The solution given in equation [5.8] is dependent on eight unknown constants for every value of m and n in equations [5.15] and [5.16]. These are Am , Bm , Cm , Dm , En , Fn , Gn and Hn . The plate has four edges, and on each of them we can prescribe two quantities; therefore, we obtain eight unknowns per boundary condition case. Edge Deflection

Rotation

Moment

Reaction

Φx (− a2 , y)

Mx (− a2 , y)

Qx (− a2 , y)

1

w(− a2 , y)

2

w(x, − 2b ) Φy (x, − 2b ) My (x, − 2b ) Qy (x, − 2b )

3

w( a2 , y)

Φx ( a2 , y)

Mx ( a2 , y)

Qx ( a2 , y)

4

w(x, 2b )

Φy (x, 2b )

My (x, 2b )

Qy (x, 2b )

Table 5.2. Possible edge conditions

It is convenient to number the edges in a counter-clockwise direction, as given in Table 5.2, for all the four edges, and from every row, i.e. the edge, only two relations can be used, depending on the particular edge restraints. The expanded expressions for the edge deflection, slope, bending moment and shear can be obtained by substituting the solution [5.8] into equations [5.2]–[5.7], which are given in section 5.7, Appendix A.

New Analytic Solutions for Elastic Buckling of Isotropic Plates

97

As an example, we derive the equation that will be obtained when we apply zero deflection along edge 2. After substitution of y = −b/2 into equation [5.8] for the deflection along the edge, we have    ∞  b ¯ m × Am cosh( α1 b ) + Bm cosh( α2 b ) w x, − = X 2 2 2 m=1 α1 b α2 b ) − Dm sinh( ) 2 2 ∞   Y¯n | y=− b × En cosh(α3 x) + Fn cosh(α4 x) + − Cm sinh(

n=1

2

+ Gn sinh(α3 x) + Hn sinh(α4 x)

[5.24]

In the second summation, we have the hyperbolic functions that are dependent on x; however, this term is multiplied by Y¯n | y=− b , which is zero along the edge. 2 Note that wp along the four edges is zero, as it is the solution for the simply supported plate (SSSS). Then, equation [5.24] becomes, as given in equation [A.4] in section 5.7, Appendix A:    ∞  b ¯ m × Am cosh( α1 b ) + Bm cosh( α2 b ) w x, − = X 2 2 2 m=1 α1 b α2 b ) − Dm sinh( ) [5.25] 2 2 ¯ m is a function of x, the edge deflection can be zero if we have Then, as X  α1 b α2 b α1 b α2 b ) + Bm cosh( ) − Cm sinh( ) − Dm sinh( ) 0 = Am cosh( 2 2 2 2 [5.26] − Cm sinh(

In this equation, we have four unknowns for a given value of m. Similarly, we can derive the equations for the remaining seven edge conditions. We will obtain a total of 4(m + n) equations. The buckling load of the plate is found when a zero force P will generate a unit defection under the applied point load (Eisenberger and Deutsch 2015). We take a unit force P = 1, and calculate the unknown constants in equations [5.15] and [5.16]. The next step is to calculate the deflection at that particular point of loading using equation [5.8], or, for better convergence, we instead find the value of (1/w(ξ, η)) that becomes zero. A bisection procedure is applied to obtain the final convergence to the desired accuracy. The buckling mode is then calculated as w(x, y) for the particular in-plane loading. As an example, the equations that are needed to solve case 48, CGSF are given in Table 5.3. This case provides a good perspective of how each equation is used, and covers all types of possible boundary conditions. In this table, we make use of the equations in section 5.7, Appendix A.

98

Modern Trends in Structural and Solid Mechanics 1

Edge # Equations 1–4

1-C W (− a2 , y)

2-G = 0 φy (x, − 2b ) = 0

[A.2] Equations 5–8

φx (− a2 , y) [A.6]

3-S

[A.8] =0

Qy (x, − 2b ) [A.16]

W ( a2 , y)

4-F = 0 My (x, 2b ) = 0

[A.1] =0

Mx ( a2 , y) [A.9]

[A.11] =0 Qy (x, 2b ) = 0 [A.15]

Table 5.3. Equations required for solving case 48, CGSF

5.5. Numerical results The new exact solutions for 51 cases are presented in Table 5.4. This table contains the buckling load Nxx for a thin isotropic plate with various aspect ratios a/b. Some cases have the same numbering, and this number represents the case number in Table 5.1; however, the first appearance is for the loading in the x direction and the second for the loading in the y direction. All the answers are compared to the finite element results from the ANSYS commercial code. The models used the SHELL181-type elements, with a mesh containing 10,000 elements (100 divisions in both directions). The results summarized in Table 5.4 reveal several important points regarding the exact solution. The Levy-type cases that are solved (cases 1–6, 22–25 and 34–35) all result in the known analytic solution, whereas the very detailed and fine mesh used in the finite element solution (approximately 30,000 degrees of freedom, depending on the particular case) did not fully converge to the exact result. The finite element results are, in all cases, above the correct values as the theory predicts, i.e. convergence from above. For several cases, an important and unexpected result is obtained: the uniaxial and biaxial buckling loads are very close. The factor of reduction given here for these cases is defined as the ratio of the uniaxial buckling load divided by the biaxial load for a square plate: in cases 6, SFSF and 24, SGSF, the ratio is 1.022, in cases 25, SGSG and 34, CGSG, it is 1.0, and in case 35, it is 1.045. In these cases, the addition of loading in the perpendicular direction has a very small effect on the buckling load (i.e. below 5%). In Table 5.5, the buckling load Nxx is presented for plates with biaxial loading. Five different cases with different aspect ratios a/b and various compression ratios β are provided. These cases are 7, CCCC, 11, CFCF, 13, CCSS, 23, SCSG and 48, CGSF. As in Table 5.4, we compare all our results with ANSYS, which are indicated by the numbers in parentheses.

New Analytic Solutions for Elastic Buckling of Isotropic Plates

Case number

1 SSSS

2 SCSS

2 CSSS

3 SFSS

3 SSFS

4 SCSC

4 CSCS

5 SFSC

a b

0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3

Uniaxial Biaxial, β = 1 Present study ANSYS Present study ANSYS 61.6850 61.6860 49.3480 49.3487 39.4784 39.4816 19.7392 19.7410 42.8368 42.8480 14.2561 14.2578 39.4784 39.4840 12.3370 12.3387 39.4784 39.4860 10.9662 10.9680 67.6377 67.6410 52.9493 52.9515 56.6536 56.6646 26.2798 26.2847 53.6038 53.6170 22.3775 22.3835 55.3250 55.3530 21.2702 21.2767 53.6038 53.6297 20.6118 20.6187 102.5104 102.5300 85.0810 85.1004 47.8394 47.8480 26.2798 26.2847 44.3814 44.3910 16.3656 16.3684 41.8144 41.8260 13.2373 13.2395 40.5468 40.5610 11.2283 11.2302 42.9897 42.9930 39.1818 39.1832 13.8332 13.8341 10.4138 10.4146 8.4657 8.4660 4.9466 4.9470 6.5943 6.5944 2.9359 2.9361 5.2618 5.2618 1.3958 1.3959 20.1630 20.1656 11.7435 11.7448 23.3497 23.3500 10.4138 10.4146 22.7479 22.7491 9.9687 9.9698 22.7999 22.8020 9.7954 9.7968 22.7986 22.8024 9.6790 9.6806 75.9099 75.9190 58.4697 58.4764 75.9099 75.9394 37.7996 37.8147 70.2314 70.2490 37.1845 37.2041 68.8070 68.8300 37.7480 37.7699 69.6322 69.6980 37.1845 37.2057 179.5024 179.6030 150.9919 151.0730 66.5526 66.5830 37.7996 37.8147 53.0460 53.0680 19.8349 19.8404 47.8394 47.8700 14.6174 14.6206 43.4900 43.5230 11.5877 11.5898 44.0507 44.0540 39.5601 39.5608 16.3096 16.3103 11.2865 11.2870 12.4127 12.7440 6.2308 6.2310 12.8463 13.1850 4.5266 4.5268 12.7438 12.7450 3.3558 3.3559

Table 5.4. Buckling load Nxx for a plate with different plate aspect ratios and boundary conditions (ν = 0.3)

99

100

Modern Trends in Structural and Solid Mechanics 1

Case number

5 CSFS

6 SFSF

6 FSFS

7 CCCC

8 CCSC

8 CCCS

9 CCCF

9 CCFC

a b

0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3

Uniaxial Biaxial, β = 1 Present study ANSYS Present study ANSYS 25.9190 25.9211 18.1064 18.1072 23.6093 23.6100 11.2865 11.2870 22.9254 22.9268 10.2134 10.2145 22.8029 22.8050 9.8900 9.8914 22.7990 22.8030 9.7008 9.7023 38.4187 38.4240 37.6026 37.6066 9.3989 9.4004 9.2005 9.2019 4.1140 4.1147 4.0467 4.0473 2.2921 2.2924 2.2646 2.2650 1.0088 1.0089 1.0020 1.0022 15.6142 15.6152 9.0583 9.0597 20.1630 20.1637 9.2005 9.2019 22.2690 22.2692 9.3184 9.3198 22.3121 22.3135 9.4007 9.4021 22.7479 22.7512 9.4977 9.4993 190.8646 190.9700 154.8911 154.9770 99.4259 99.4872 52.3447 52.3721 82.4162 82.4790 40.6749 40.6975 77.6449 77.7230 38.7228 38.7456 72.6338 72.7003 38.1018 38.1262 114.5872 114.6200 92.1244 92.1489 79.8152 79.8475 42.5471 42.5645 72.6338 72.6640 38.1018 38.1213 71.1388 71.1770 37.7621 37.7833 69.9107 69.9510 37.3599 37.3817 184.3691 184.4730 151.0482 151.1250 79.6206 79.6600 42.5471 42.5645 69.1642 69.2190 27.0099 27.0193 61.4150 61.4520 23.0311 23.0388 57.4940 57.5332 21.0408 21.0480 161.4800 161.6500 111.8760 112.4730 45.1631 45.1877 28.5408 28.5628 24.8789 24.8850 12.7745 12.7749 19.0019 19.0040 7.4374 7.4380 16.6207 16.6240 4.1766 4.1768 41.7247 41.7296 29.7479 29.7482 38.5512 38.5920 28.5408 28.5628 38.1469 38.2918 28.0731 28.1481 37.9796 38.2713 27.9690 28.1226 37.5601 38.2785 27.7300 28.1255

Table 5.4. (cont.) Buckling load Nxx for a plate with different plate aspect ratios and boundary conditions (ν = 0.3)

New Analytic Solutions for Elastic Buckling of Isotropic Plates

Case number

10 CSCF

10 SCFC

11 CFCF

11 FCFC

12 CFSF

12 FCFS

13 CCSS

14 CCSF

a b

0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3

Uniaxial Biaxial, β = 1 Present study ANSYS Present study ANSYS 160.6445 160.8100 111.8755 112.4730 43.1486 43.1727 28.2924 28.3140 21.4604 21.4670 12.6880 12.6883 13.8897 13.8910 6.7676 6.7696 8.4920 8.4938 2.6323 2.6326 41.6190 41.6246 27.0710 27.0753 37.9941 38.0334 28.2924 28.3140 38.1353 38.2798 28.0233 28.0980 37.9789 38.2706 27.9689 28.1225 37.5600 38.2786 27.7300 28.1225 156.0736 156.2300 111.1686 111.7310 38.6773 38.7011 27.0569 27.0782 17.0604 17.0690 10.5295 10.5302 9.5363 9.5405 5.3367 5.3371 4.1942 4.1963 2.0860 2.0861 26.2953 26.2975 21.3469 21.3472 36.0469 36.0802 27.0654 27.0798 37.6781 37.8180 27.5693 27.6042 37.7440 38.0385 27.8668 27.9370 37.5502 38.2662 27.9844 28.1041 79.3608 79.3990 72.9186 73.0130 19.5658 19.5723 17.6330 17.6379 8.5993 8.6020 7.5490 7.5493 4.7962 4.7976 4.1161 4.1164 2.1068 2.1074 1.7521 1.7521 20.0705 20.0725 16.4639 16.4648 30.1965 30.2113 17.6330 17.6380 29.0975 29.1390 18.0726 18.0869 30.1447 30.2285 18.2297 18.2564 29.9862 30.2385 18.2623 18.3323 107.6543 107.6800 87.6908 87.7112 61.4150 61.4310 32.0524 32.0603 57.9389 57.9610 24.0495 24.0563 55.9436 55.9630 21.9227 21.9295 54.5604 54.5870 20.7802 20.7872 84.8584 84.8990 73.2299 73.3222 26.1719 26.1780 18.4733 18.4782 16.6136 16.6150 8.7530 8.7532 14.5176 14.5180 5.6383 5.6388 13.7841 13.7862 3.6813 3.6813

Table 5.4. (cont.) Buckling load Nxx for a plate with different plate aspect ratios and boundary conditions (ν = 0.3)

101

102

Modern Trends in Structural and Solid Mechanics 1

Case number

14 CCFS

15 CSSF

15 SCFS

22 SSSG

22 SSGS

23 SCSG

23 CSGS

24 SGSF

a b

0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3

Uniaxial Biaxial, β = 1 Present study ANSYS Present study ANSYS 32.6935 32.6948 22.5533 22.5552 31.0005 31.0136 18.4733 18.4782 30.2031 30.2502 18.3219 18.3353 30.1446 30.2414 18.3075 18.3337 29.9957 30.2397 18.2651 18.3336 83.9445 83.9830 73.2285 73.3210 24.0286 24.0347 18.2502 18.2553 12.9830 12.9850 7.9912 7.9920 9.1295 9.1298 4.4092 4.4097 6.3842 6.3846 1.8871 1.8872 30.2075 30.2137 17.6358 17.6359 30.2273 30.2390 18.2502 18.2553 30.1856 30.2327 18.3172 18.3305 30.1393 30.2359 18.3071 18.3334 29.9956 30.2397 18.2650 18.3334 44.5674 44.5720 41.9458 41.9502 15.4213 15.4224 12.3370 12.3382 10.7092 10.7100 6.8539 6.8545 9.8696 9.8698 4.9348 4.9352 10.7092 10.7115 3.5640 3.5643 39.4784 39.4795 19.7392 19.7402 46.3323 46.3410 12.3370 12.3376 39.4784 39.4830 10.9662 10.9673 41.4770 41.4910 10.4865 10.4878 40.4240 40.4440 10.1438 10.1463 45.9395 45.9440 42.9320 42.9364 18.9775 18.9789 14.6174 14.6190 17.5579 17.5590 10.3843 10.3854 18.9775 18.9830 9.4499 9.4510 17.5579 17.5581 9.2961 9.2975 66.5526 66.5631 37.7996 37.8044 52.0625 52.0760 14.6174 14.6190 43.4900 43.5010 11.5877 11.5892 42.1055 42.1160 10.7330 10.7346 40.6060 40.6180 10.2118 10.2134 38.9000 38.9060 38.1984 38.2025 9.6047 9.6062 9.4007 9.4020 4.2199 4.2206 4.1260 4.1266 2.3497 2.3501 2.3001 2.3005 1.0285 1.0287 1.0117 1.0118

Table 5.4. (cont.) Buckling load Nxx for a plate with different plate aspect ratios and boundary conditions (ν = 0.3)

New Analytic Solutions for Elastic Buckling of Isotropic Plates

Case number

24 GSFS

25 SGSG

25 GSGS

26 CSCG

26 SCGC

27 CSSG

27 SCGS

34 CGSG

a b

0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3

Uniaxial Biaxial, β = 1 Present study ANSYS Present study ANSYS 25.5927 25.5946 9.2005 9.2018 22.3121 22.3124 9.4007 9.4020 22.8514 22.8528 9.4977 9.4992 22.7977 22.7998 9.5496 9.5511 22.7989 22.8027 9.5989 9.6005 39.4784 39.4840 39.4784 39.4844 9.8696 9.8712 9.8696 9.8712 4.3865 4.3872 4.3865 4.3872 2.4674 2.4678 2.4674 2.4678 1.0966 1.0968 1.0966 1.0968 61.6850 61.6935 9.8696 9.8724 39.4784 39.4840 9.8696 9.8725 42.8368 42.8490 9.8696 9.8715 39.4784 39.4850 9.8696 9.8700 39.4784 39.4860 9.8696 9.8712 162.8865 163.0600 148.2573 148.4030 44.8694 44.9015 37.7418 37.7700 23.5197 23.5310 16.4823 16.4894 16.6403 16.6450 9.4522 9.4537 13.2622 13.2656 4.9599 4.9600 84.9562 84.9613 37.8089 37.8154 68.7958 68.8260 37.7418 37.7702 69.6109 69.6600 37.7874 37.8175 70.6640 70.7410 37.0643 37.1027 68.7517 68.8360 37.1252 37.1605 85.8036 85.8490 81.6383 81.6827 25.6261 25.6340 21.2687 21.2762 15.0340 15.0370 10.2804 10.2826 11.9603 11.9620 6.5706 6.5712 11.0954 11.0970 4.0918 4.0919 56.6637 56.6640 26.2824 26.2836 55.3216 55.3370 21.2687 21.2762 56.6413 56.6640 20.6084 20.6184 53.3804 53.4070 20.4096 20.4213 53.7118 53.7430 20.2820 20.2928 80.7438 80.7890 80.7437 80.7895 20.1893 20.1980 20.1893 20.1977 8.9740 8.9768 8.9740 8.9768 5.0484 5.0494 5.0484 5.0495 2.2438 2.2442 2.2438 2.2442

Table 5.4. (cont.) Buckling load Nxx for a plate with different plate aspect ratios and boundary conditions (ν = 0.3)

103

104

Modern Trends in Structural and Solid Mechanics 1

Case number

34 GCGS

35 CGCG

35 GCGC

43 CCCG

43 CCGC

44 CCSG

44 CCGS

47 CGCF

a b

0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3 0.5 1 1.5 2 3

Uniaxial Biaxial, β = 1 Present study ANSYS Present study ANSYS 67.6472 67.6493 20.1937 20.1979 56.6419 56.6670 20.1893 20.1977 53.5871 53.6180 20.1874 20.1977 55.3025 55.3530 20.1859 20.1978 53.5664 53.6300 20.1869 20.1977 157.8384 158.0100 150.8336 151.0750 39.4725 39.5040 37.7879 37.8160 17.5471 17.5570 17.5471 17.5570 9.8723 9.8761 9.8723 9.8761 4.3881 4.3894 4.3881 4.3894 75.9174 75.9288 39.4894 39.5040 75.8948 75.9420 37.7879 37.8158 70.1767 70.2500 37.1574 37.2044 68.7359 68.8310 37.7084 37.7700 69.5338 69.6980 37.1552 37.2057 164.0203 164.2000 151.0096 151.2320 47.7090 47.7430 38.7141 38.7440 28.7828 28.7950 18.9092 18.9163 24.8573 24.8670 13.0888 13.0910 20.6080 20.6160 10.1697 10.1710 99.4369 99.4625 52.3551 52.3620 77.6258 77.6770 38.7141 38.7439 73.6483 73.7350 38.3798 38.4270 71.2663 71.3720 37.7524 37.8094 69.7470 69.8970 37.3843 37.4695 87.0036 87.0510 82.2652 82.3140 28.6451 28.6540 23.0291 23.0373 20.6042 20.6080 13.2907 13.2932 19.9539 19.9580 10.6374 10.6387 18.1588 18.1630 9.5256 9.5270 79.6332 79.6390 42.5495 42.5535 65.0655 65.0980 23.0291 23.0373 57.4752 57.5120 21.0364 21.0474 55.6955 55.7430 20.5663 20.5791 54.3515 54.4150 20.3214 20.3328 156.8678 157.0100 111.8619 112.4580 39.0308 39.0580 27.9114 27.9335 17.2669 17.2770 12.3124 12.3142 9.6709 9.6753 7.3331 7.3331 4.2664 4.2673 3.9171 3.9177

Table 5.4. (cont.) Buckling load Nxx for a plate with different plate aspect ratios and boundary conditions (ν = 0.3)

New Analytic Solutions for Elastic Buckling of Isotropic Plates

Case number

a b

0.5 1 47 GCFC 1.5 2 3 0.5 1 48 CGSF 1.5 2 3 0.5 1 48 GCFS 1.5 2 3

105

Uniaxial Biaxial, β = 1 Present study ANSYS Present study ANSYS 36.0835 36.0838 29.3306 29.3328 38.4653 38.5065 27.9114 27.9335 38.1114 38.2555 28.0653 28.1403 37.9793 38.2710 27.9655 28.1189 37.5599 38.2785 27.7299 28.1254 79.9855 80.0240 73.2185 73.3111 19.8428 19.8500 18.4050 18.4101 8.7549 8.7576 8.2860 8.2875 4.8919 4.8931 4.7017 4.7025 2.1502 2.1505 2.1003 2.1008 30.2152 30.2175 18.8046 18.8093 30.2116 30.2246 18.4050 18.4101 30.1885 30.2353 18.3192 18.3327 30.1389 30.2355 18.3046 18.3310 29.9955 30.2397 18.2649 18.3334

Table 5.4. (cont.) Buckling load Nxx for a plate with different plate aspect ratios and boundary conditions (ν = 0.3)

In Figure 5.1, the first buckling mode shapes for three cases are shown. These buckling modes are for four different aspect ratios a/b for cases 27, SCGS, 34, GCGS and 44, CCGS, all with uniaxial loading in the x direction. Here we observe that a change in the plate aspect ratio results in a very significant change in the buckling mode. Figure 5.2 shows the modes for the same cases as Figure 5.1, but for biaxial loading with β = 1. Unlike for the unidirectional loading case in Figure 5.1, we observe here that the buckling modes remain unchanged when the aspect ratio is increased. 5.6. Conclusion The analytical solution for the problem of isotropic plate buckling with many of the possible combinations of boundary conditions is given. This new solution does not require the separation between symmetric and antisymmetric cases; instead, they can be obtained using only one expression. Splitting the variables in the solution allows us to represent the edge conditions for deflection, slope, shear forces and bending moment in terms of common functions. Finally, when we apply the eight particular boundary conditions to the edges, we can find the set of equations that are required to determine all the terms in the complete solution. The presented results can be used as benchmark results for further studies on this topic.

106

Modern Trends in Structural and Solid Mechanics 1

Case

a b

0.5 1 7 CCCC 1.5 2 0.5 1 11 CFCF 1.5 2 0.5 1 13 CCSS 1.5 2 0.5 1 23 SCSG 1.5 2 0.5 1 48 CGSF 1.5 2

-1 240.3182 (240.4700) 147.7043 (147.8400) 126.3699 (126.5200) 118.0669 (118.2200) 156.7479 (156.8600) 38.9199 (38.9440) 17.1819 (17.1920) 9.6036 (9.6089) 136.9266 (136.9600) 105.8388 (105.8900) 98.6346 (98.6880) 95.8631 (95.9180) 49.3099 (49.3140) 26.6181 (26.6190) 29.9064 (29.9160) 26.6181 (26.6190) 80.2532 (80.2910) 19.9495 (19.9570) 8.8166 (8.8195) 4.9322 (4.9335)

-0.5 213.3413 (213.4700) 129.4421 (129.5700) 107.6367 (107.7300) 97.9599 (98.0460) 156.5588 (156.6800) 38.8404 (38.8650) 17.1373 (17.1470) 9.5764 (9.5817) 120.7793 (120.8100) 87.6193 (87.6560) 78.0992 (78.1260) 75.4577 (75.4890) 47.5731 (47.5770) 22.1972 (22.1980) 26.3233 (26.3230) 22.1972 (22.2010) 80.1639 (80.2020) 19.9116 (19.9190) 8.7939 (8.7967) 4.9167 (4.9180)

β 0 190.8646 (190.9700) 99.4259 (99.4872) 82.4162 (82.4790) 77.6449 (77.7230) 156.0736 (156.2300) 38.6773 (38.7011) 17.0604 (17.0690) 9.5363 (9.5405) 107.6543 (107.6800) 61.4150 (61.4310) 57.9389 (57.9610) 55.9436 (55.9630) 45.9395 (45.9440) 18.9775 (18.9789) 17.5579 (17.5590) 18.9775 (18.9830) 79.9855 (80.0240) 19.8428 (19.8500) 8.7549 (8.7576) 4.8919 (4.8931)

0.5 171.7551 (171.8500) 69.3919 (69.4290) 61.2038 (61.2400) 56.8042 (56.8420) 150.5267 (150.9100) 37.2665 (37.2900) 16.2680 (16.2700) 8.9530 (8.9542) 96.8168 (96.8400) 42.5475 (42.5580) 36.6602 (36.6700) 36.2018 (36.2130) 44.3968 (44.4010) 16.5332 (16.5350) 13.0772 (13.0780) 13.1155 (13.1170) 79.3718 (79.4130) 19.6597 (19.6670) 8.6666 (8.6692) 4.8431 (4.8443)

1 154.8911 (154.9770) 52.3447 (52.3721) 40.6749 (40.6975) 38.7228 (38.7456) 111.1686 (111.7310) 27.0569 (27.0782) 10.5295 (10.5302) 5.3367 (5.3371) 87.6908 (87.7112) 32.0524 (32.0603) 24.0495 (24.0563) 21.9227 (21.9295) 42.9320 (42.9364) 14.6174 (14.6190) 10.3843 (10.3854) 9.4499 (9.4510) 73.2185 (73.3111) 18.4050 (18.4101) 8.2860 (8.2875) 4.7017 (4.7025)

Table 5.5. First buckling load Nxx for rectangular plates with different aspect ratios a/b and different compression ratios β (ν =0.3)

New Analytic Solutions for Elastic Buckling of Isotropic Plates

Figure 5.1. Buckling modes for plates with different aspect ratios a/b under uniaxial loading. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

107

108

Modern Trends in Structural and Solid Mechanics 1

Figure 5.2. Buckling modes for plates with different aspect ratios a/b under biaxial loading. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

New Analytic Solutions for Elastic Buckling of Isotropic Plates

109

5.7. Appendix A: Deflection, slopes, bending moments and shears In this appendix, we present 16 possible edge boundary conditions for the plate. a. The four displacement equations along the edges of the plate ∞  α a α a  a   ¯ 3 4 + Fn cosh Yn × En cosh W x = ,y = 2 2 2 n=1 α a  α a  3 4 + Hn sinh +Gn sinh 2 2 ∞   α a α a a   ¯ 3 4 W x = − ,y = + Fn cosh Yn × En cosh 2 2 2 n=1 α a  α a  3 4 − Hn sinh −Gn sinh 2 2

[A.1]

[A.2]

    

  ∞ b α1 b α2 b ¯ W x, y = = + Bm cosh Xm × Am cosh 2 2 2 m=1     α1 b α2 b + Dm sinh [A.3] +Cm sinh 2 2

W

    

  ∞ b ¯ m × Am cosh α1 b + Bm cosh α2 b x, y = − = X 2 2 2 m=1     α1 b α2 b − Dm sinh [A.4] −Cm sinh 2 2

b. The four slope equations normal to the edges of the plate  φx

∞  α a α a a   ¯ 3 4 + Fn α4 sinh x = ,y = Yn × En α3 sinh 2 2 2 n=1 α a  α a  3 4 + Hn α4 cosh +Gn α3 cosh 2 2 ∞  λa cos(mπ) × (Am cosh (α1 y) + m=1

+Bm cosh (α2 y) + Cm sinh (α1 y) + Dm sinh (α2 y)) +

∞ ∞   m=1 n=1

4 P sin (ξ λa ) sin (η λb ) λa cos(mπ) Y¯n × dmn ab [A.5]

110

Modern Trends in Structural and Solid Mechanics 1

 φx

∞  α a α a a   ¯ 3 4 − Fn α4 sinh x = − ,y = Yn × −En α3 sinh 2 2 2 n=1 α a  α a  3 4 + Hn α4 cosh +Gn α3 cosh 2 2 ∞  + λa × (Am cosh (α1 y) + Bm cosh (α2 y) m=1

+Cm sinh (α1 y) + Dm sinh (α2 y)) +

∞ ∞   m=1 n=1

4 P sin (ξ λa ) sin (η λb ) λa Y¯n × dmn ab

[A.6]

    

  ∞ b ¯ m × Am α1 sinh α1 b + Bm α2 sinh α2 b φy x, y = = X 2 2 2 m=1     α1 b α2 b + Dm α2 cosh +Cm α1 cosh 2 2 +

∞ 

λb cos(nπ) × (En cosh (α3 x)

n=1

+Fn cosh (α4 x) + Gn sinh (α3 x) + Hn sinh (α4 x)) +

∞ ∞   m=1 n=1

¯ m × 4 P sin (ξ λa ) sin (η λb ) λb cos(nπ) X dmn ab [A.7]

 φy

b x, y = − 2





 α1 b ¯ = − Xm × −Am α1 sinh 2 m=1     α2 b α1 b + Cm α1 cosh + Bm α2 sinh 2 2   α2 b Dm α2 cosh 2 ∞ 

+

∞ 

λb × (En cosh (α3 x) + Fn cosh (α4 x)

n=1

+Gn sinh (α3 x) + Hn sinh (α4 x)) +

∞ ∞   m=1 n=1

¯ m × 4 P sin (ξ λa ) sin (η λb ) λb X dmn ab

[A.8]

New Analytic Solutions for Elastic Buckling of Isotropic Plates

111

c. The four bending moment equations along the edges of the plate ∞    α a a   ¯ 3 Mx x = , y = Yn D × − En α3 2 cosh 2 2 n=1 α a α a 4 3 + Gn α3 2 sinh +Fn α4 2 cosh 2 2  α a  4 2 +Hn α4 sinh 2  α a α a 3 4 + Fn cosh + ν λb 2 × En cosh 2 2 α a  α a  3 4 + Hn sinh +Gn sinh 2 2

[A.9]

∞    α a a   ¯ 3 Mx x = − , y = Yn D × − En α3 2 cosh 2 2 n=1 α a 4 +Fn α4 2 cosh 2 α a  α a  3 4 2 − Hn α4 2 sinh −Gn α3 sinh 2 2  α a α a 3 4 2 + Fn cosh + ν λb × En cosh 2 2 α a  α a  3 4 − Hn sinh [A.10] −Gn sinh 2 2

 My

b x, y = 2

 =





 α1 b − Am α1 2 cosh 2 m=1   α2 b +Bm α2 2 cosh 2     α1 b α2 b + Dm α2 2 sinh +Cm α1 2 sinh 2 2  

  α1 b α2 b 2 + Bm cosh + ν λa × Am cosh 2 2     α1 b α2 b + Dm sinh [A.11] +Cm sinh 2 2 ∞ 

¯mD × X

112

Modern Trends in Structural and Solid Mechanics 1

 My

b x, y = − 2

 =

∞ 

¯ mD × X

m=1

+Bm α2 2 cosh



2



− Am α1 cosh 

α2 b 2

α1 b 2





 α2 b − Dm α2 sinh −Cm α1 sinh 2  

  α1 b α2 b 2 + Bm cosh + ν λa × Am cosh 2 2

    α1 b α2 b − Dm sinh [A.12] −Cm sinh 2 2 2



α1 b 2



2



d. The four reaction equations along the edges of the plate ∞  a   Qx x = , y = Dλ3a cos(mπ) × (Am cosh (α1 y) 2 m=1

+Bm cosh (α2 y) + Cm sinh (α1 y) + Dm sinh (α2 y))  α a 3 + D Y¯n × En α3 3 sinh 2 n=1 α a 4 Fn α4 3 sinh 2 α a  α a  3 4 + Hn α4 3 cosh +Gn α3 3 cosh 2 2  ∞   + D(ν − 2) λa cos(mπ) × Am α1 2 cosh (α1 y)



∞ 

m=1 2

+ Bm α2 cosh (α2 y) + Cm α1 2 sinh (α1 y)  + Dm α2 2 sinh (α2 y)  α a α a 3 4 + Fn α4 sinh λ2b Y¯n × En α3 sinh 2 2 n=1  α a  α a  3 4 + Hn α4 cosh +Gn α3 cosh 2 2



∞ 

New Analytic Solutions for Elastic Buckling of Isotropic Plates

 + Nxx

∞ 

113

λa cos(mπ) × (Am cosh (α1 y)

m=1

+Bm cosh (α2 y) + Cm sinh (α1 y)  α a 3 Y¯n × En α3 sinh 2 n=1 α a α a 4 3 + Gn α3 cosh +Fn α4 sinh 2 2   α a  4 +Hn α4 cosh 2

+Dm sinh (α2 y)) +

∞ 

∞ ∞  

4P sin (ξ λa ) sin (η λb ) λa cos(mπ) Y¯n × dmn ab m=1 n=1   × D λb 2 ν − D λa 2 − 2 D λb 2 − Nxx [A.13] −

∞  a   Dλ3a × (Am cosh (α1 y) + Bm cosh (α2 y) Qx x = − , y = 2 m=1

+Cm sinh (α1 y) + Dm sinh (α2 y))  α a 3 D Y¯n × −En α3 3 sinh 2 n=1 α a 4 −Fn α4 3 sinh 2 α a  α a  3 4 + Hn α4 3 cosh +Gn α3 3 cosh 2 2  ∞   + D(ν − 2) λa × Am α1 2 cosh (α1 y)



∞ 

m=1 2

+ Bm α2 cosh (α2 y) + Cm α1 2 sinh (α1 y)  + Dm α2 2 sinh (α2 y)  α a 3 λ2b Y¯n × −En α3 sinh 2 n=1 α a α a 4 3 + Gn α3 cosh −Fn α4 sinh 2 2   α a  4 +Hn α4 cosh 2



∞ 

114

Modern Trends in Structural and Solid Mechanics 1

 + Nxx

∞ 

λa (Am cosh (α1 y) + Bm cosh (α2 y)

m=1

+Cm sinh (α1 y)  α a 3 Y¯n × −En α3 sinh 2 n=1 α a α a 4 3 + Gn α3 cosh −Fn α4 sinh 2 2   α a  4 +Hn α4 cosh 2

+Dm sinh (α2 y)) +

∞ 

∞ ∞  

4P sin (ξ λa ) sin (η λb ) λa Y¯n × dmn ab m=1 n=1   × D λb 2 ν − D λa 2 − 2 D λb 2 − Nxx −

[A.14]

     ∞  b ¯ m × Am α1 3 sinh α1 b =− Qy x, y = DX 2 2 m=1   α2 b +Bm α2 3 sinh 2     α1 b α2 b 3 3 + Dm α2 cosh +Cm α1 cosh 2 2 −

∞ 

Dλ3b cos(nπ) × (En cosh (α3 x)

n=1

+Fn cosh (α4 x) + Gn sinh (α3 x) + Hn sinh (α4 x))  

 ∞  α1 b 2 ¯ λa Xm × Am α1 sinh + D(ν − 2) − 2 m=1     α2 b α1 b + Cm α1 cosh +Bm α2 sinh 2 2   α2 b +Dm α2 cosh 2 +

∞  n=1

 λb cos(nπ) × En α3 2 cosh (α3 x)

New Analytic Solutions for Elastic Buckling of Isotropic Plates

+Fn α4 2 cosh (α4 x) 2

2

+Gn α3 sinh (α3 x) + Hn α4 sinh (α4 x) 



115





 ¯ m × Am α1 sinh α1 b + X 2 m=1   α2 b Bm α2 sinh 2     α1 b α2 b + Dm α2 cosh +Cm α1 cosh 2 2

+ Nyy

+

∞ 

∞ 

λb cos(nπ) × (En cosh (α3 x)

n=1



+Fn cosh (α4 x) + Gn sinh (α3 x) + Hn sinh (α4 x)) ∞ ∞  

¯ m × 4P sin (ξ λa ) sin (η λb ) λb cos(nπ) X dmn ab m=1 n=1   × D λa 2 ν − D λb 2 − 2 D λa 2 − Nyy [A.15] −

     ∞  b ¯ m × −Am α1 3 sinh α1 b =− DX Qy x, y = − 2 2 m=1   α2 b −Bm α2 3 sinh 2     α1 b α2 b 3 3 + Dm α2 cosh +Cm α1 cosh 2 2 +

∞ 

Dλ3b × (En cosh (α3 x)

n=1

+Fn cosh (α4 x) + Gn sinh (α3 x) + Hn sinh (α4 x))  

 ∞  α1 b 2 ¯ λa Xm × −Am α1 sinh + D(ν − 2) − 2 m=1     α2 b α1 b + Cm α1 cosh −Bm α2 sinh 2 2   α2 b +Dm α2 cosh 2

116

Modern Trends in Structural and Solid Mechanics 1

+

∞ 

 λb × En α3 2 cosh (α3 x) + Fn α4 2 cosh (α4 x)

n=1

 +Gn α3 sinh (α3 x) + Hn α4 sinh (α4 x) 2





2



 α1 b ¯ Xm × −Am α1 sinh + Nyy 2 m=1   α2 b −Bm α2 sinh 2     α1 b α2 b + Dm α2 cosh +Cm α1 cosh 2 2 +

∞ 

∞ 

λb × (En cosh (α3 x) + Fn cosh (α4 x)

n=1



+Gn sinh (α3 x) + Hn sinh (α4 x)) ∞  ∞ 

¯ m × 4P sin (ξ λa ) sin (η λb ) λb X dmn ab m=1 n=1   × D λa 2 ν − D λb 2 − 2 D λa 2 − Nyy



[A.16]

5.8. Appendix B: Function transformation In this appendix, the expansions of equations [5.15] and [5.16] are derived. By rewriting Xn , we obtain Xn = En cosh (α3 x) + Fn cosh (α4 x) + Gn sinh (α3 x) + Hn sinh (α4 x) [B.1] We are interested in expanding the four cosh and sinh hyperbolic functions in ¯ m , which are as follows: terms of the trigonometric functions in X cosh(αi x) =

∞ 

Rm sin

 mπx 

m=2,4...

+

∞  m=1,3...

Sm cos

a

cos

 mπx  a

 mπ 

sin

2  mπ  2

[B.2]

New Analytic Solutions for Elastic Buckling of Isotropic Plates

117

with Rm

2 = a

Sm

2 = a



a 2 −a 2



a 2 −a 2

cosh(αi x)sin

cosh(αi x)cos

 mπx  a

 mπx  a

cos

sin

 mπ  2

 mπ  2

dx = 0

dx =

[B.3]

4 π m cosh( α2i a ) [B.4] m2 π 2 + α2i a2

and using trigonometric identities, we finally arrive at the following, simpler, expression: cosh(αi x) =

∞ 

sin

m=1,2,3...

 mπ 2 4 π m cosh( αi a ) 2 ¯m, X 2 m2 π 2 + α2i a2

[B.5]

Similarly, we obtain, for sinh(αi x) sinh(αi x) =

∞ 

−cos

m=1,2,3...

 mπ 2 4 π m sinh( αi a ) 2 ¯ m, X 2 m2 π 2 + α2i a2

[B.6]

and cosh(αi y) =

∞ 

sin

n=1,2,3...

sinh(αi y) =

∞ 

 nπ 2 4 π n cosh( αi b ) 2 Y¯n , 2 n2 π 2 + α2i b2

−cos

n=1,2,3...

 nπ 2 4 π n sinh( αi b ) 2 Y¯n , 2 n2 π 2 + α2i b2

[B.7]

[B.8]

Here we can rewrite the expressions for the hyperbolic functions sinh and cosh in a compact form as sinh(α1 y) =

∞ 

s1 Y¯n ,

[B.9]

s2 Y¯n ,

[B.10]

¯m, s3 X

[B.11]

n=1

sinh(α2 y) =

∞  n=1

sinh(α3 x) =

∞  m=1

118

Modern Trends in Structural and Solid Mechanics 1

sinh(α4 x) =

∞ 

¯m, s4 X

[B.12]

c1 Y¯n ,

[B.13]

c2 Y¯n ,

[B.14]

¯m, c3 X

[B.15]

¯m, c4 X

[B.16]

m=1

cosh(α1 y) =

∞  n=1

cosh(α2 y) =

∞  n=1

cosh(α3 x) =

∞  m=1

cosh(α4 x) =

∞  m=1

s1 = −4 π n

α1 b cos2 ( nπ 2 ) sinh( 2 ) , n2 π 2 + α1 2 b2

[B.17]

s2 = −4 π n

α2 b cos2 ( nπ 2 ) sinh( 2 ) , n2 π 2 + α2 2 b2

[B.18]

s3 = −4 π m

α3 a cos2 ( mπ 2 ) sinh( 2 ) , m2 π 2 + α3 2 a2

[B.19]

s4 = −4 π m

α4 a cos2 ( mπ 2 ) sinh( 2 ) , m2 π 2 + α4 2 a2

[B.20]

c1 = 4 π n

α1 b sin2 ( nπ 2 ) cosh( 2 ) , n2 π 2 + α1 2 b2

[B.21]

c2 = 4 π n

α2 b sin2 ( nπ 2 ) cosh( 2 ) , n2 π 2 + α2 2 b2

[B.22]

c3 = 4 π m

α3 a sin2 ( mπ 2 ) cosh( 2 ) , 2 2 2 m π + α3 a2

[B.23]

c4 = 4 π m

α4 a sin2 ( mπ 2 ) cosh( 2 ) . m2 π 2 + α4 2 a2

[B.24]

New Analytic Solutions for Elastic Buckling of Isotropic Plates

119

5.9. References Bert, C.W. and Malik, M. (1994). Frequency equations and modes of free vibrations of rectangular plates with various edge conditions. Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science, 208, 308–319. Bulson, P.S. (1970). The Stability of Flat Plates. Chatto & Windus, London. Dickinson, S. (1978). The buckling and frequency of flexural vibration of rectangular isotropic and orthotropic plates using Rayleigh’s method. Journal of Sound and Vibration, 61(1), 1–8. Eisenberger, M. and Deutsch, A. (2015). Static analysis for exact vibration analysis of clamped plates. International Journal of Structural Stability and Dynamics, 15, 1540030. Levy, M. (1899). Sur l’´equilibre e´ lastique d’une plaque rectangulaire. Comptes rendus de l’Acad´emie des Sciences de Paris, 129(1), 535–539. Li, R., Zheng, X., Wang, H., Xiong, S., Yan, K., Li, P. (2018). New analytic buckling solutions of rectangular thin plates with all edges free. International Journal of Mechanical Sciences, 144, 67–73. Liu, X., Liu, X., Xie, S. (2020). A highly accurate analytical spectral flexibility formulation for buckling and wrinkling of orthotropic rectangular plates. International Journal of Mechanical Sciences, 168, 105311. Narita, Y. and Leissa, A.W. (1990). Buckling studies for simply supported symmetrically laminated rectangular plates. International Journal of Mechanical Sciences, 32(11), 909–924. ´ Navier, L. (1819). R´esum´e des lec¸ons de m´ecanique. Ecole Polytechnique, Paris. Reddy, J.N. (1997). Mechanics of Laminated Composite Plates: Theory and Analysis. CRC Press, Boca Raton. Shufrin, I. and Eisenberger, M. (2005). Stability and vibration of shear deformable plates – First order and higher order analyses. International Journal of Solids and Structures, 42(3), 1225–1251. Shufrin, I., Rabinovitch, O., Eisenberger, M. (2008a). Buckling of laminated plates with general boundary conditions under combined compression, tension, and shear – A semi-analytical solution. Thin-Walled Structures, 46(7), 925–938. Shufrin, I., Rabinovitch, O., Eisenberger, M. (2008b). Buckling of symmetrically laminated rectangular plates with general boundary conditions – A semi-analytical approach. Composite Structures, 82(4), 521–531. Szilard, R. (2004). Theories and Application of Plate Analysis. John Wiley & Sons, New York. Timoshenko, S.P. (1959). Theory of Plates and Shells. McGraw-Hill, Reading. Timoshenko, S.P. and Gere, J.M. (1961). Theory of Elastic Stability. McGraw-Hill, New York. Ullah, S., Zhang, J., Zhong, Y. (2019a). Accurate buckling analysis of rectangular thin plates by double finite sine integral transform method. Structural Engineering and Mechanics, 72, 491–502. Ullah, S., Zhong, Y., Zhang, J. (2019b). Analytical buckling solutions of rectangular thin plates by straightforward generalized integral transform method. International Journal of Mechanical Sciences, 152, 535–544.

120

Modern Trends in Structural and Solid Mechanics 1

Wang, C.M., Wang, C.Y., Reddy, J.N. (2005). Exact Solutions for Buckling of Structural Members. CRC Press, Boca Raton. Wang, B., Li, P., Li, R. (2016). Symplectic superposition method for new analytic buckling solutions of rectangular thin plates. International Journal of Mechanical Sciences, 119, 432–441.

6 Buckling and Post-Buckling of Parabolic Arches with Local Damage

6.1. Introduction The birth of fracture mechanics is attributed to Griffith (1920), who based it on the solutions provided by Kirsch (1898) and Inglis (1913). Many efforts to enhance this new branch of solid mechanics followed (Irwin 1957a,b, 1960a,b; Dugdale 1960; Wells 1963; Rice 1968; Rice and Levy 1972). Applications of fracture mechanics to one-dimensional continua were possible after the introduction of the stress intensity factor (SIF) in Irwin (1957a,b). This, together with Castigliano’s theorem, allows crack-like damages to be modeled by springs, and the response of beam frames can be described with no need to find the stress field around the crack tip. The structural response includes the displacement profiles due to benchmark load distributions, as well as the frequencies and mode shapes of natural free vibration, which encompass information about geometry, material properties and boundary conditions. A decade before the introduction of SIFs, Kirshmer (1944) and Thomson (1949) focused on simple structures with crack-type damages using a reduced area. Even after the introduction of SIFs, these crack models were used by many researchers. Dimarogonas (1981, 1982, 1996); Dimarogonas and Paipetis (1983); Anifantis and Dimarogonas (1983a,b,c) successfully used the spring analogy by means of SIFs, especially on rotating machines. These papers paved the way to many more, which focus on building enhanced models for either the crack or the structure and on examining the effects of cracks on the static and dynamic behavior of beam-type structures (Gounaris and Dimarogonas 1988; Ostachowicz and Krawwczuk 1992;

˘ Chapter written by U˘gurcan E RO GLU , Ekrem T ÜFEKCI.

Giuseppe RUTA,

Achille PAOLONE and

Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

122

Modern Trends in Structural and Solid Mechanics 1

Chati et al. 1997; Khiem and Lien 2001; Sinha and Friswell 2002; Viola et al. 2002; Krawczuk et al. 2003; Kisa and Arif Gurel 2007; Orhan 2007; Mazanoglu et al. 2009; Caddemi and Morassi 2013; Caddemi et al. 2017; Caliò et al. 2017; Cannizzaro et al. 2017; Eroglu and Tufekci 2017b; Eroglu et al. 2019, 2020), without pretending to be exhaustive. A thorough investigation of the structural response suggested that monitoring the mechanical behavior of cracked elements may lead to damage identification by suitable methods. Thus, many papers focused on experimental settings, optimization techniques and their effects on identification procedures (Cawley and Adams 1979; Ju and Minovich 1986; Cerri and Ruta 2004; Cerri et al. 2008; Pau et al. 2010; Ciambella and Vestroni 2015; Capecchi et al. 2016; Eroglu and Tufekci 2016, 2017a; Caddemi and Morassi 2007, 2011; Greco and Pau 2011). A survey of the entire literature is out of our scope, yet, we may refer to the review of Doebling et al. (1996) for details. Most studies on cracked one-dimensional structural elements deal with their statics and free dynamics, while their stability is only given marginal consideration, especially arches. A possible reason for this is that stability analysis is usually related to structural design and functioning, and is not meant for structural monitoring purposes. We may quote Karaagac et al. (2011), who examined the static and dynamic stability of circular arches by the finite element technique. They used the SIFs provided in Müller et al. (1993) and assumed the crack on the inner surface of the arch. We investigate buckling and post-buckling of parabolic arches with crack-like damages. The latter may be simplified models of either the failure of connections of large structures and actual cracks in relatively small-size applications. We model the arch as a one-dimensional curved beam and study its fundamental and bifurcated path under reasonable kinematic and constitutive assumptions. The crack is modeled by the spring analogy and the SIFs, following literature and previous works by the authors. The novel points here are: a) the fundamental path is non-trivial, in that we linearize the finite field equations and find a non-zero linear elastic response to a uniform load with respect to the arch span; b) the critical values of the load multiplier are found in closed form; c) the germ of the post-buckling path is found by an additional perturbation of the finite field equations about the bifurcation point, providing information on the quality of the bifurcated path. Thorough comments and mechanical interpretations, in light of classic results on stability, are provided, together with hints on possible applications and future developments. 6.2. A one-dimensional model for arches We model arches as initially curved fully deformable beams, i.e. one-dimensional structured continua that are assumed flexible, extensible and shear-deformable. Then, we suppose that their reference shape is well defined starting from a segment of a

Buckling and Post-Buckling of Parabolic Arches with Local Damage

123

regular planar curve, called the arch axis. The curve segment lies on a plane. On the latter we fix a Cartesian frame x, y, to which a consistent basis of unit vectors {ex , ey } is associated. A third unit vector ez , which coincides with one of the unit normals to the plane of the arch axis, completes a vector basis for the vector space associated with the three-dimensional Euclidean ambient space. The position of any point P of the axis is described by the vector field r0 r0 (x) = xex + y(x)ey ,

0≤x≤x ¯, y(0) ≤ y ≤ y(¯ x)

[6.1]

The tangent vector to the arch axis at any point P and its unit counterpart l(s) are  dy(x) dr0 (x) dr0 (x) dr0 (x) = ex − ey , · dx, ds = dx dx dx dx [6.2] dr0 (x(s)) dx(s) l(x(s)) = , dx(s) ds where ds is the arc length, which is a measure of an element of the axis compatible with the given Cartesian metrics. Equation [6.2]2 yields the intrinsic abscissa s in terms of x and vice versa; hence, r0 and all the fields depending on P may be expressed in terms of either x or s; thus, equation [6.2]3 gives the field of unit vectors tangent to the axis. The s-derivative of the field l(x(s)), its magnitude and its unit counterpart are dl(x(s)) dx(s) dl(x(s)) = =: k(x(s))m(x(s)), ds dx(s) ds 1 dl(x(s)) m(x(s)) = k(x(s)) ds

[6.3]

The quantity k is the local curvature of the axis at P ; the Frénet–Serret local basis consists of the triad {l, m, n}, n = l × m. Since the axis is a plane curve, the unit tangents and normals depend on P , while the bi-normal does not and coincides, with the exception of the a sign depending on the location of the osculating circle at P , with ez . The reference shape is a three-dimensional region that is completed by attaching the copies of a prototype plane figure to any point of the axis. These are called transverse cross-sections and, being all equal, model an arch with uniform geometry with respect to the axis. With no loss in generality, the cross-sections are attached to the axis orthogonally to the unit tangent to the axis. That is, the initial setting of the cross-sections is determined by the Frénet–Serret local triad of unit base vectors for any point of the axis: the arch element is along the unit tangent l, the corresponding cross-section lies in the plane spanned by the unit normal m and the unit bi-normal n. For simplicity, the dependence on x(s) will be dropped if no confusion arises.

124

Modern Trends in Structural and Solid Mechanics 1

6.2.1. Finite kinematics and balance, linear elastic law The reference configuration is supposed to be stress-free (i.e. the arch is in an initial natural state). A different shape is provided by the fields d, the vector of the axis displacement (equivalently, by the new position r = r0 + d ∀ P ), and R, the orthogonal tensor of the cross-section’s rotation; both are functions of P , hence of x (or s) and of an evolution parameter (time, for instance). Neutral, or rigid, changes of shape are characterized by a uniform value of R and by a tangent vector in the new shape that is simply the R-transformed of l. If primes denote s-derivatives to shorten the notation, finite deformation measures in the actual configuration are thus (see Antman (1973, 1995) and Pignataro et al. (2008)) 

v = (r0 + d) − Rl,

V = R R

[6.4]

The components of the vector field v are the elongation ε of the axis and the shearing strains γ between the axis and the cross-sections; the non-zero components of the skew-symmetric tensor field V are the variation of curvature χ of the axis. If we suppose the loads and the change of shape take place in the same plane of the referential configuration, R is expressed in terms of a single angle ϑ, providing the rotation of the cross-section at P about the bi-normal n, and the displacement field has just two components:   cos ϑ − sin ϑ , d = ul + vm [6.5] (R) = sin ϑ cos ϑ Then, we have only two and one components of the strain measures v, V, respectively ε = 1 − cos ϑ + kv + u ,

γ = − sin ϑ + ku + v  ,

χ = ϑ

[6.6]

The environment acts on the arch by: a vector force field, power dual of the incremental displacement of the axis; and a skew-symmetric tensor couple field, power dual of the incremental rotation of the cross-sections. Force and couple may be either distributed (actions at a distance) and localized at the end cross-sections (actions by contact); they are denoted b, B, f , F, respectively. The parts of the arch interact by contact via a vector and a skew symmetric tensor field, power duals of the incremental strains, denoted t, T, respectively. All the action fields are functions of the place P along the axis and of the evolution parameter, in general. The balance equations in the configuration at a given value of the evolution parameter, also called actual shape, may be derived either by applying the virtual

Buckling and Post-Buckling of Parabolic Arches with Local Damage

125

work principle or by assuming balance of force and torque as a basic principle. In both cases, they read (see Antman (1973, 1995) and Pignataro et al. (2008)) t + b = 0,



T + (r0 + d) ∧ t + B = 0 ∀x ∈ (−l, l), t = ±t±l ,

T = ±T±l

x = ±l

[6.7]

with ∧ the external product of vectors, providing skew-symmetric tensors. The balance of torque has a parallel expression in terms of axial vectors of skew-symmetric tensors and the cross-product between them. The two components of the inner force and the sole component of the inner couple, with respect to the local basis in the actual ˜ , T˜, M ˜ , respectively; they are usually named normal configuration, will be denoted N and transverse (or shearing) force and bending couple, respectively. The components of the inner actions are resultants and resultant moments of Cauchy stresses (i.e., surface densities of contact force in the actual shape). It is customary in structural mechanics, however, to express them in terms of the inner actions in the reference configuration, with components denoted N, T, M , respectively (this is what is usually done by Cauchy and Piola stresses in three-dimensional continuum mechanics). The two sets of components are related via the rotation tensor R in equation [6.5]1 , ruling the change of unit vectors solidal to the cross-sections. Then, the scalar consequences of equation [6.7] in the presence of planar external actions with components n, t, m, with respect to the local basis in the actual shape, read N  − kT + n = 0,

T  + kN + t = 0,

M  − N (sin ϑ + γ) + T (cos ϑ + ε) + m = 0

[6.8]

Here, the local curvature k appears due to the Frénet–Serret formulas for the s-derivatives of the intrinsic local basis. Even though kinematics is in a finite setting and balance is written in the actual configuration in terms of the inner actions in the reference shape, we admit that the inner actions are linearly related to the strain measure components by elastic uncoupled constitutive laws, according to ˜ = EAε, N

T˜ = GAs γ,

˜ = EIχ M

[6.9]

where E, G are Young’s and transverse elastic moduli of the material of the arch, respectively, and A, As , I are the cross-sectional area, shearing area and second moment of area relative to a principal axis of inertia parallel to n, respectively. The positions in equation [6.9] are reasonable in that we will operate a perturbation of the finite field equations, and linear (incremental) constitutive equations suffice to

126

Modern Trends in Structural and Solid Mechanics 1

describe the arch response. The uncoupled laws in equation [6.9], similar to those provided by Saint-Venant’s theory, are reasonable enough for compact cross-sections with main dimensions that are small compared with the local radius of the osculating circle (see Timoshenko and Goodier (1951)). Then, taking into account that the local frame attached to the cross-section changes according to the orthogonal tensor R in equation [6.5], the constitutive relations in equation [6.9] can be given in terms of the components of the inner action in the reference shape, providing  2    cos ϑ sin2 ϑ 1 1 ε=N + − + T sin ϑ cos ϑ , EA GAs EA GAs     2 1 1 M cos ϑ sin2 ϑ + N sin ϑ cos ϑ − , χ= + γ=T GAs EA EA GAs EI [6.10] In the following sections, we describe how we use the set of field equations for the arch in order to follow its response to an external load uniformly distributed with respect to the metrics of the span and monotonically increasing. A perturbation approach will follow a linearized solution in the neighborhood of the referential configuration (fundamental path), until a critical value of the load multiplier is reached. This corresponds to a bifurcation of the static equilibrium; in order to follow the jet of the bifurcated path, a second perturbation of the field equations in the neighborhood of the buckled shape will be performed. 6.2.2. Non-trivial fundamental equilibrium path The field equations [6.6], [6.8] and [6.10] form a system of ordinary differential equations with respect to the spatial coordinate of any point P of the arch axis. To find the static response to any set of distributed external actions n, t, m in the actual shape, we are supposed to solve this nonlinear differential system, supplemented by the relevant boundary conditions. Let all external actions be rescaled by a load multiplier: then, the solutions of the field equations provide a one-parameter family of new shapes, called the fundamental path, which is described in exact, nonlinear, finite form. It goes without saying that, bar very special cases, the resolution of the field equations in finite form is computationally very demanding. However, in structural engineering, we usually do not face finite displacements, and we may assume the structural response to be limited to “small” displacements and rotations. Hence, linearization of the field equations is possible, and the search for the fundamental path is greatly simplified. To this aim, let us introduce an evolution parameter η ∈ [0, 1] such that η = 0, η = 1 characterize the reference and the actual shape, respectively. A perturbation approach, i.e., a formal power series expansion

Buckling and Post-Buckling of Parabolic Arches with Local Damage

127

of all the fields of interest in terms of η, provides an approximate response of the structure about a given value of the evolution parameter (see Pignataro et al. (1991)). If we thus let η = 0 and limit the formal expansion at the first order, we study a nontrivial neighborhood of the reference configuration, supposed to represent the natural state of the arch. Then, if the suffix 0 denotes quantities evaluated in the reference shape, R0 = I ⇔ ϑ0 = 0,

d0 = 0,

N0 = T 0 = M 0 = 0

[6.11]

Hence, the linear expansion of equations [6.6], [6.8] and [6.10] about the unstressed reference state provides EA(u˙ f − k v˙ f ) = N˙ f , N˙ f − k T˙f + n˙ f = 0,

GAs (v˙ f − ϑ˙ f + k u˙ f ) = T˙f , T˙f + k N˙ f + t˙f = 0,

EI ϑ˙ f = M˙ f ,

M˙ f + T˙f + m ˙ f = 0,

[6.12]

where the overdot denotes η-derivatives of functions evaluated at η = 0; the suffix f indicates fields along the fundamental path. As is well known, the fundamental path is a single-valued curve expressing a characteristic displacement (or strain) versus the load multiplier when the latter monotonically grows from zero. When the load multiplier attains a so-called critical value, the equilibrium path becomes a multivalued curve and bifurcation of the static solution occurs (see Pignataro et al. (1991)). 6.2.3. Bifurcated path To investigate the static bifurcation phenomenon in a neighborhood of the critical point, we assume that any function g of interest in this problem may be written in the additive form g = gf + gb , where the subscripts f and b denote the values of g in the fundamental and bifurcated paths, respectively. If we let gb = 0, we investigate the fundamental path only; if gb = 0 and we add it to gf , we obtain the value attained by g in the bifurcated path when departing from the fundamental one. Let us investigate a neighborhood of the bifurcation point, found by evaluating the critical value of the load multiplier along the fundamental path. Then, we suppose that any function of interest along the bifurcated path regularly depends on another evolution parameter β, equal to zero at the bifurcation point and growing along the bifurcated path. We may thus perform a power series expansion of gb in terms of β about the critical point β = 0 (see Pignataro et al. (1991)); if overdots now stand for the β-derivatives evaluated at β = 0 and gb (β = 0) = 0, i.e. g(β = 0) = gf , we have g + ... gb = 0 + β g˙ + (β 2 /2)¨

[6.13]

Performing such an expansion for the system of equations [6.6], [6.8] and [6.10], where all functions of interest are now evaluated along the bifurcated path, we obtain

128

Modern Trends in Structural and Solid Mechanics 1

a hierarchy of systems of ordinary differential equations, one at each order of the β-formal power series expansion (see Pignataro et al. (1991)). The first set of the hierarchy provides the first-order germ of the bifurcated path and accounts for the quantities of the fundamental path at the critical point   1 1 N˙ b  ˙ −k v˙ b + sin ϑf ϑb + u˙ b = + Tf ϑb − = ε˙b , EA EA GAs   1 1 T˙b − = γ˙ b , + Nf ϑb k u˙ b − cos ϑf ϑ˙ b + v˙ b = GAs EA GAs [6.14] M˙ b , N˙ b − k T˙b + n˙ b = 0, T˙b + k N˙ b + t˙b = 0, ϑ˙ b = EI M˙ b − Nf (γb + cos ϑf ϑ˙ b ) − N˙ b (γf + sin ϑf )+ +Tf (ε˙b − sin ϑf ϑ˙ b ) + T˙b (cos ϑf +εf ) + m ˙b=0 The terms along the fundamental path may be replaced with those evaluated by the η-linearization discussed in the previous section, since we assumed that the fundamental path, however non-trivial, consists of shapes adjacent to the reference configuration. Thus, equation [6.14] becomes   1 1 Nb + Tf ϑb − − ϑf ϑb − kvb = EA EA GAs   1 1 Tb  − , vb + kub − ϑb = + Nf ϑb GAs EA GAs EIϑb = Mb Nb −kTb +nb = 0, Tb +kNb + tb = 0, ub

[6.15]

Mb −Nf (γb +ϑb )−Nb (γf +ϑf )+Tf (εb −ϑf ϑb )+Tb (1+εf )+mb = 0 where all overdots are omitted for a simpler notation: all quantities are thus actually first-order increments (with the exception of the curvature k in the reference configuration): those with subscript f are with respect to the reference configuration, and those with subscript b with respect to the critical state. The solution to the set in equation [6.15] provides the first-order bifurcated path with respect to β, accounting for the strains and inner actions attained at the bifurcation point. Next, we show a couple of meaningful benchmark examples to show how the proposed two-parameter perturbation expansions work and provide known results. 6.2.4. Special benchmark examples Here, we provide qualitative evidence on the generality of the differential system [6.15]. We present two benchmark examples from the literature; since shearing strain is neglected in these, we admit the same inner constraint, without loss of generality.

Buckling and Post-Buckling of Parabolic Arches with Local Damage

129

Straight compressed beams. In a straight (k = 0) and slender (γ = 0) beam subjected only to a compressive force P , it is immediate to see that the non-trivial linear response, describing the fundamental path, is γf = 0,

εf = −P/EA,

ϑf = 0,

Nf = −P,

Tf = Mf = 0

[6.16]

Substituting equation [6.16] into equation [6.15] leads to the set of linear differential equations EAub = Nb ,

vb = ϑb (1 − εf ),

Nb = Tb = 0,

EIϑb = Mb ,

Mb + P ϑb + Tb (1 − εf ) = 0

[6.17]

which can be rearranged as EAub = 0,

EIvbIV + P (1 − εf ) vb = 0

[6.18]

Equations [6.18] are identical to the first two of (31) in Pignataro and Ruta (2003), and we also find them in Pflüger (1964). If we neglect the initial contraction, i.e. εf = 0, equations [6.18] coincide with the well-known differential equations for Euler buckling. Circular arch under radial load. A portion of a slender (γ = 0) circular arch of radius R has uniform initial curvature k = 1/R and is loaded by a uniform radial distribution of magnitude q with the same direction throughout the motion (i.e. “dead” or conservative). If, as is usual in the literature, the arch is supported by rollers, the only non-zero fields of the fundamental path follow “Mariotte’s formula” εf = −(qR)/(EA),

Nf = −qR

[6.19]

Substituting equation [6.19] into equation [6.15] leads to the set of linear differential equations   Nb ub qR vb   ub − = , vb − ϑb 1 − + = 0, R EA EA R Mb Tb , Nb − − qϑb = 0, ϑb = [6.20] EI R   qR Nb = 0, Mb + qRϑb + Tb 1 − =0 Tb + R EA If the axial strain is negligible, i.e. 1 − (qR/EA) ≈ 1, the system [6.20] becomes      vb  ub u  2  IV b EA ub − − qR vb + − REI vb + =0 R  R   R  [6.21]    u v ub EI vb + b + REA ub − b − qR vb + =0 R R R

130

Modern Trends in Structural and Solid Mechanics 1

Equation [6.21] is identical to those of Simitses and Hodges (2006), after suitably changing the independent variable and the orientation of the radial displacement. From these two benchmark studies, we deduce that the two-parameter perturbation expansions of the finite field equations proposed here are a reliable tool for investigating the static stability of non-trivial equilibria. In the next section, we investigate parabolic arches, which are the main objectives of our study. 6.3. Parabolic arches Parabolic arches are curved beams, the axes of which are described by a segment of parabola, symmetric with respect to its vertex. We fix the coordinate frame as in Figure 6.1; thus, the position vector of the points of the axis is r0 (x) = xex + f [1 − (x2 /l2 )]ey ,

−l ≤ x ≤ l, 0 ≤ y ≤ f

[6.22]

in terms of the Cartesian abscissa x along the span, 2l long; f is the height of the arch crown with respect to the line of the span, connecting the two ends. y

sP P f

ey

r0

l

m

ex xP

O

x

2l

Figure 6.1. Reference shape of a parabolic arch with cross-sections orthogonal to the axis

By equations [6.2] and [6.3], the tangent vectors, their magnitude (the arc length), their unit counterpart l, the local curvature and unit normals are  2x 4f 2 x2 l2 ex − 2f xey dr0 dx, l =  , = ex − f 2 ey , ds = 1 + 4 dx l l l4 + 4f 2 x2 [6.23] 2f l4 2f xex + l2 ey k= , m = − 3/2 l4 + 4f 2 x2 (l4 + 4f 2 x2 )

Buckling and Post-Buckling of Parabolic Arches with Local Damage

131

The Frénet–Serret triad in equation [6.23] is completed by the bi-normal n = l × m = −ez (the sign is due to the assigned curvature, with osculating circle below the axis ∀P ). To abstract from the particular values of the geometrical and physical quantities in a problem, it is customary to operate non-dimensional analyses; let us then define  (x, s, u, v) A f GAs (¯ x, s¯, u ¯, v¯) = , α= , λ=l , A¯ = , l l I EA [6.24] (N, T )l2 Ml (n, t)l3 ml2 ¯ ¯ ¯ ¯ (N , T ) = , M= , (¯ n , t) = , m ¯ = EI EI EI EI Equation [6.23]2 provides (at least locally) the functions x(s) and s(x); thus, we may also deduce that x¯ = x¯(¯ s) and equation [6.24] provides further non-dimensional quantities  ds d¯ s = = L(¯ x) = 1 + 4α2 x ¯2 , dx d¯ x

¯ x) = kl = k(¯

2α . L3 (¯ x)

[6.25]

Substituting equations [6.24] and [6.25] into equation [6.12], the equations for the non-dimensional η-first-order fundamental path are written in matrix form dyf = Af yf +qf , d¯ x

 ¯f T¯f M ¯f T , yf = u ¯f v¯f ϑf N ⎛

1 0 λ2 1 1 0 ¯ 2 Aλ 0 0 0 0 0 k 0 −k 0 0 0 −1

0 k0

⎜ ⎜ ⎜−k ⎜ Af = L ⎜ ⎜0 ⎜0 ⎜ ⎝0

T qf = −L{0 0 0 n ¯ f t¯f m ¯ f} ,

0

0 0 0 0 0

⎞ 0 ⎟ ⎟ 0⎟ ⎟ 1⎟ ⎟. 0⎟ ⎟ 0⎠ 0

[6.26]

A force load −qey that is uniform with respect to the span length implies that mf = 0 ⇒ m ¯ f = 0; its components with respect to the arc length in the local basis are given by the force equivalence −qey dx = (nl + tm)ds

[6.27]

whence their non-dimensional counterparts according to equations [6.24]. The solution of equation [6.26] is (Hubbard and West 1995; Tufekci et al. 2017)    x¯ −1 x) = Yf (¯ x) yf (¯ x0 ) + Yf (ξ) q (ξ) dξ [6.28] yf (¯ 0

132

Modern Trends in Structural and Solid Mechanics 1

where Y(¯ x) is called the principal matrix, or matricant (Pease 1965), or transfer matrix (Pestel and Leckie 1963), of the homogeneous equation [6.26] about x¯ = 0. The entries of the principal matrix for a planar curved beam in integral form can be found in Tufekci and Arpaci (2006). They are provided in Eroglu et al. (2020) for parabolic arches of uniform cross-sections; variable cross-sections are considered in Eroglu and Ruta (2020). The state vector y in equation [6.28] accounts for distributed external loads; concentrated ones at any point x ¯i ∈ [−1, 1] were recently discussed in Eroglu et al. (2020) through continuity of displacement and local balance of actions. We admit that the given load is “dead”; then, it is easy to prove that nb = −tf ϑb , tb = nf ϑb . Substituting equations [6.24] and [6.25] into equations [6.15] provides the equations for the non-dimensional β-first-order bifurcated path in matrix form dyb d¯ x ⎛ ⎜ 0 ⎜ ⎜ ⎜ −k ⎜ ⎜ Ab = L ⎜ ⎜ 0 ⎜ ⎜ 0 ⎜ ⎜ 0 ⎝ 0

 ¯b T¯b M ¯b T yb = u ¯b v¯b ϑb N

= Ab yb , k −ϑf + Tf (

1 λ2

0

0

0 0 0

1 1 − ¯ 2) 2 λ Aλ 1 1 1 + Nf ( 2 − ¯ 2 ) λ Aλ 0 tf −nf

0

Nf

0 0 −k Tf + ϑf ¯ Aλ2

0 1 ¯ 2 Aλ 0 k 0 Nf −1 − ¯ 2 Aλ

⎞ 0⎟ ⎟ ⎟ 0⎟ ⎟ [6.29] ⎟ ⎟ 1⎟ ⎟ 0⎟ ⎟ 0⎟ ⎠ 0

In a homogeneous arch, the mass per unit length and the bending stiffness are uniform, but the terms in equation [6.29] depend on x ¯, and Ab cannot be reduced to an upper triangular form. Hence, equation [6.29], in general, does not have closed-form solutions. Thus, Eroglu et al. (2019) propose to find approximate solutions by the Peano series (Peano 1888) and Volterra’s multiplicative integral (Pease 1965; Slavik 2007) yb (¯ x) = Yb (¯ x, x ¯0 )yb (¯ x0 ) Yb (¯ x, x ¯0 ) =

n  ı=1

Y2 (¯ x0 +ıΔ¯ x, x ¯0 +(ı−1)Δ¯ x),

x2 , x¯1 ) ≈ I + Ab (¯ x1 )(¯ x2 − x ¯1 )+ Y2 (¯     1 d2 Ab  1 2 1 dAb  + + A (¯ x1 ) (¯ x2 − x ¯1 )2 + 2 dx x¯1 4 dx2 x¯1 2 b

[6.30]

x = (¯ x − x¯0 ) /n. As n increases, Δ¯ x→0 having split {¯ x, x ¯0 } in n intervals with Δ¯ and equation [6.30] turns into Volterra’s integral (Slavik 2007). We keep n finite, yet large enough to ensure convergence.

Buckling and Post-Buckling of Parabolic Arches with Local Damage

133

Equation [6.30]1 requires a point x ¯0 where the values of the displacements and contact actions are known. This occurs at the boundary, where non-dimensional natural conditions are (see equation [6.24]): clamped end : u¯ = 0, v¯ = 0, ϑ = 0, ¯ = 0, pinned end : u¯ = 0, v¯ = 0, M ¯ = 0, T¯ = 0, M ¯ = 0. free end : N

[6.31]

They are three such natural conditions at each end, yielding six linear equations in the six components yb T (q) yb (¯ x0 ) = 0

[6.32]

with T(q) a square 6 × 6 matrix of coefficients. Non-trivial solutions yb (¯ x0 ) require the singularity of T(q) in terms of the load multiplier q, yielding a highly nonlinear equation, the solutions of which are the critical loads for the arch, det [T (q)] = 0 ⇒ q = qcr

[6.33]

6.4. Crack models for one-dimensional elements If the arch has a small plane crack on the cross-section at x ¯c , its damaging effect is described by considering two regular arch chunks connected at xc by a set of springs; as in Eroglu and Tufekci (2017a), their compliances are related to the depth of the crack through its strain energy Uc , given by Tada et al. (2000)   2  2 1  ∂ 2 Uc , ı, j = N, T, M dA, cıj = + Uc = K K Ij IIj  ∂ı∂j Ac E [6.34] where Ac is the cross-section as reduced by the crack; E  = E/(1 − ν 2 ), with ν Poisson’s ratio of the material; KIj , KIIj are the stress intensity factors in opening and shearing modes, respectively, due to the j-th contact action; and cıj is the compliance of the spring representing the effect of the crack on the j-th contact action due to a unit value of the ı-th kinematic descriptor. In a rectangular cross-section of height h with a crack of depth a at opposite sides, with respect to the center of curvature, consider the additional non-dimensional quantities related to the crack a ¯ ıj = K√ıj , U ¯c = Uc , c¯N¯ N¯ = EI cN N , , K h hEA l3 E h EI EI EI cMM , c¯M¯ N¯ = c¯N¯ M¯ = 2 cMN = 3 cT T , c¯M¯ M¯ = l l l

a ¯= c¯T¯T¯

[6.35]

134

Modern Trends in Structural and Solid Mechanics 1

Following Tada et al. (2000), the shape functions fi of the crack magnitude ratio  2 tan(π¯ a/2) 0.752 + 2.02¯ a + 0.37[1 − sin(π¯ a/2)]3 (π¯ a/2), f1 (¯ a) = π¯ a cos a3 1.22 − 0.561¯ a + 0.085¯ a2 + 0.18¯ √ , 1−a ¯  4 2 tan(π¯ a/2) 0.923 + 0.199 [1 − sin (π¯ a/2)] a) = f3 (¯ π¯ a cos (π¯ a/2)

f2 (¯ a) =

[6.36] and equation [6.34] yield the non-dimensional compliances  a¯ √ ξ 2 f (ξ) dξ, c¯N¯ N¯ = 4π 3(1 − ν 2 ) 3 1 0 λ √ 1 − ν  a¯ ξ c¯T¯T¯ = 16π 3 f 2 (ξ)dξ, 1 + ν 0 A¯2 λ3 2  a¯ √ ξ 2 f3 (ξ) dξ, c¯M¯ M¯ = 12π 3(1 − ν 2 ) 0 λ  a¯ ξ c¯N¯ M¯ = 12π(1 − ν 2 ) f (ξ) f3 (ξ) dξ. 2 1 0 λ

[6.37]

Henceforth, only non-dimensional quantities will be used; with another abuse of notation, overbars will be omitted for simplicity, except when confusion may arise. The crack at xc implies a jump of the axis displacement in the actual shape; the jump depends on the contact actions at the cross-sections of the left and right chunks of the arch (whence the superscripts l, r) facing each other at xc . The inner contact actions at the ends of the fictitious springs will, of course, be balanced. These vector conditions are written in the local basis of the left chunk, pulled by the operator R: ˜r − d ˜ l = C ∗˜ R (ϑr − ϑl ) d f l,

R (ϑr − ϑl ) ˜ fr = ˜ fl

[6.38]



⎞ cN N 0 pcN M C∗ = ⎝ 0 c T T 0 ⎠ pcN M 0 cMM The tilde here denotes vectors in actual configuration. For easier calculations, inner actions f = {N , T , M }T and displacement are written with respect to their components in the reference frame; hence, equation [6.38] becomes R(ϑr −ϑl )RT (ϑr )dr −RT (ϑl )dl = C∗ RT (ϑl )f l , R(ϑr −ϑl )RT (ϑr )f r = RT (ϑl )f l

[6.39]

Buckling and Post-Buckling of Parabolic Arches with Local Damage

135

Since R (ϑr − ϑl ) RT (ϑr ) = RT (ϑl ), equation [6.39] may be written in matrix form yr (xc ) = C(a, p)yl (xc ),   I R (ϑl (xc )) C∗ RT (ϑl (xc )) , C (a, p) = 0 I

[6.40]

where 0, I are the 3 × 3 null and identity matrices, respectively, and p = ±1 is a non-material parameter indicating that the crack is at the top or bottom of the cross-section with respect to the center of curvature, respectively. Equation [6.40] is nonlinear and can be subjected to the same perturbation performed for field equations. When it is linearized about the initial unstressed configuration, the condition yr (xc ) = Cf yl (xc ), identical to that in Eroglu et al. (2020), is recovered. A linearization about the non-trivial fundamental path leads to the matrix C depending on ϑb , rearranged as yr (xc ) = Cb yl (xc )

[6.41]

where Cb is too long to be reported and will be omitted for the sake of space. For both right and left chunks, equation [6.28], solution for the static problem, holds; the unknown initial values are now 12, and adding the jump to the boundary conditions yields 12 equations for calculating them. It is then better to present the jump conditions in terms of these initial values; if ˆI is the 6 × 6 identity, equation [6.40] becomes     xc −1 −1 r l ˆ yf (x0 ) = Yf (xc ) Cf Yf (xc )yf (x0 )+ Cf − I Yf (ξ) q (ξ) dξ 0

[6.42] For the eigenvalue problem equation [6.33] we will then update the principal matrix for a given crack location xc inside the j-th interval between x0 and x as follows: Yb (x, x0 ) = Yb (x, x0 + jΔx) Ybc Yb (x0 + (j − 1) Δx, x0 ) Ybc = Y2 (x0 + jΔx, xc ) Cb Y2 (xc , x0 + (j − 1) Δx)

[6.43]

6.5. An application Consider a fixed parabolic arch with geometric aspect α = 0.4 and slenderness ratio ς = 100, under the described load; if the cross-section is rectangular, A¯ = 0.3. The arch is assumed to have a transverse crack at x = xc with depth ratio a = 0.6

136

Modern Trends in Structural and Solid Mechanics 1

(recall that, to simplify the notation, we dropped the overbars for the non-dimensional quantities related to the crack). We refer to Eroglu et al. (2020) for a detailed examination of the static problem providing the first-order strain increments, which are assumed to be a sufficiently accurate representation of the actual deformed shape. If the crack is not at the crown, one of the arch chunks is more “slender” than the other. Then, it is possible that when the former buckles, the latter goes on along its fundamental path; thus, two different principal matrices should be used. In general, both chunks buckle in correspondence to the same load multiplier; in any case, we consider two different scenarios: both chunks bifurcate for the same load multiplier (“global buckling”), or the more slender reaches bifurcation, while the other remains on its fundamental path (“local buckling”). However, recall that the two sub-arches are linked by relatively stiff springs, not by a smooth hinge; i.e. the crack affects a very small portion of the cross-section and does not actually cut the arch in two.

Load ratio 1.2

     

1.1 1.0 



   

 

0.9 0.0

       

 

 

       

              

0.2

0.4



Global, p=-1



Local, p=-1

0.6  

             

           

0.8

xc 1.0

Global, p=1 Local, p=1

Figure 6.2. Critical load ratio for a fixed arch with a crack characterized by a = 0.6

Figure 6.2 presents the plot of the critical load ratio (critical load of the cracked arch divided by that of the intact arch) for an arch damaged by a crack with aspect ratio a = 0.6 versus the crack location xc . The plot considers one half of the arch,

Buckling and Post-Buckling of Parabolic Arches with Local Damage

137

because of symmetry, i.e. we plot what happens when the crack is located at the right half of the considered arch. We present the results for both “global” (square markers) and “local” buckling (round markers) and for both crack locations at the opposite sides of the cross-section with respect to the center of curvature (i.e. for the non-material parameter p = 1 (full lines) and p = −1 (dashed lines)). We see that “local” buckling requires a greater critical load than the “global” one, except for xc ≥≈ 0.8, i.e. in the vicinity of the right constraint. This has a clear physical interpretation: when the crack is near the fixed end, the difference in slenderness of the two sub-arches reaches its top; thus, the more slender chunk attains buckling before the more stubby sub-arch. The critical load ratio of this “local” buckling is very high at the beginning, then decreases as xc approaches the constrained end, resembling that of “global” buckling. This also has a physical interpretation, since for xc ≈ 0.8, the stubby chunk actually acts as an extended clamp, then makes the system globally stiffer; on the other hand, when xc → 1, the system is practically equivalent to the entire arch. Actually, for other values of xc , the global buckling prevails, which is also physically reasonable: if both chunks have comparable slenderness, they are similarly prone to buckling and bifurcate as one. We also remark that the effect of the crack position on opposite sides of the cross-section has limited effect on critical load, due to considerations similar to those provided in Eroglu et al. (2019). For more shallow arches, we may expect a stronger dependence of the critical load on p. We remark that at xc ≈ 0 and xc ≈ 0.75, the critical load ratio attains values greater than unity, i.e. the critical load multiplier of the damaged arch is greater than that of the intact one. This is a seemingly paradoxical behavior, since it was highlighted by Timoshenko and Gere (1961) that a parabolic arch with one hinge at the crown, in general, buckles under a lower critical load with respect to an entire fixed arch. However, we may check that this phenomenon occurs for cracks located at places around which the incremental bending moment vanishes. Therefore, the coupling between axial force and bending moment (accounted for by the parameter p) creates a bending moment, which acts against the buckling strain, providing a stiffening effect. Note that for the problems considered herein, the mode shape is still skew-symmetric, albeit the crack induces some slight deviations from it (see Figure 6.3). Figure 6.3 shows the buckling modes about the non-trivial fundamental path. The deformed configuration is plotted with an amplification factor of 10, to better reflect the effect of crack location on the section. Effects of crack position on the bifurcated paths are clearly shown; moreover, the question is raised about the stability of bifurcated paths and its dependence on the crack location on the section, which needs to be investigated in future contributions.

138

Modern Trends in Structural and Solid Mechanics 1

xc = 0, p = -1

xc = 0, p = 1

xc = 0.5, p = -1

xc = 0.5, p = 1 Undeformed

Fundamental path

Bifurcated path

Figure 6.3. Buckling modes about the deformed shape for a = 0.6

6.5.1. A comparison For comparison, we consider a symmetric circular arch, consisting of two regular identical chunks connected with a torsional spring with stiffness kT , as in Wang et al. (2005). This shape approximates that of a parabolic arch when its opening angle is moderate. The arch has radius R and opening angle ϕ. This is a primitive model of possible damage at the apex, neglecting axial, and shear compliances as well as axial-bending coupling. For this simplified model, in our treatment, we have a single non-zero component in the compliance matrix, cMM , which equals 1/kT . ϕ kT R EI 0 2 4 6

π/6

π/6

π/3

π/3

π/2

π/2

2π/3

2π/3

W

Present

W

Present

W

Present

W

Present

108.36∗ 129.48∗ 147.61 163.21

108.21∗ 129.30∗ 147.40 162.97

27.077∗ 36.794 43.979 49.366

27.052∗ 36.737 43.909 49.282

12.025∗ 17.956 21.751 24.277

12.008∗ 17.926 21.712 24.231

6.7578∗ 10.806 13.060 14.418

6.755∗ 10.788 13.035 14.388

Table 6.1. Critical loads for the symmetric mode shapes of a hinged circular arch with a weakened section at the apex. Stars denote that the symmetric mode has a lower critical load than the corresponding skew-symmetric mode; W=Wang et al. (2005)

In the exact treatment in Wang et al. (2005), the arch is assumed to be inextensible and shear rigid. Table 6.1 lists the critical loads for symmetric mode shapes: indeed, note that for skew-symmetric mode shape, the bending couple at the apex vanishes and the torsion spring does not affect the behavior. The weakened section only affects the critical load for symmetric mode shapes. 6.6. Final remarks We presented two perturbation expansions of the finite governing equations for arches modeled as curved beams, in order to investigate their non-trivial fundamental

Buckling and Post-Buckling of Parabolic Arches with Local Damage

139

path and their post-buckling path. Kinematics is finite, balance is in the actual configuration and only constitutive equations are supposed linear elastic. The perturbation approach lets us investigate the effect of a deformed and pre-stressed shape on the buckling load and the relevant equilibria following possible static bifurcation. We showed that this approach recovers well-known benchmark results and applied it to the analysis of buckling scenarios of parabolic arches under a “vertical” load uniformly distributed with respect to the span, affected by small cracks. The effect of the latter is suitably modeled as a set of linear elastic springs, following previous investigations; the solution of the governing equations was found numerically by using the so-called principal matrix of the differential system. We obtained results for both “local” and “global” buckling and commented on them, highlighting how seemingly paradoxical results actually have a clear physical interpretation. Future investigations will be about the quality of the post-buckling path and on linear vibration about non-trivial pre-stressed states, in order to detect the effect of local damages for monitoring purposes. 6.7. Acknowledgments This work began when U˘gurcan Ero˘glu was a visiting research student at the Dipartimento di Ingegneria Strutturale e Geotecnica of the Sapienza University of Rome, the support of which is gratefully acknowledged. Giuseppe Ruta acknowledges the support of institutional grants from Sapienza University of Rome for the year 2019. 6.8. References Anifantis, N. and Dimarogonas, A.D. (1983a). Buckling of rings and tubes with longitudinal cracks. Mechanics Research Communications, 18, 693–702. Anifantis, N. and Dimarogonas, A.D. (1983b). Post buckling behavior of transverse cracked columns. Composite Structures, 12(2), 351–356. Anifantis, N. and Dimarogonas, A.D. (1983c). Stability of columns with a single crack subjected to follower and vertical loads. International Journal of Solids and Structures, 19, 281–291. Antman, S. (1973). The Theory of Rods. Springer, Berlin. Antman, S. (1995). Nonlinear Problems of Elasticity. Springer-Verlag, New York. Caddemi, S. and Morassi, A. (2007). Crack detection in elastic beams by static measurements. International Journal of Solids and Structures, 44, 5301–5315. Caddemi, S. and Morassi, A. (2011). Detecting multiple open cracks in elastic beams by static tests. Journal of Engineering Mechanics, 137, 113–124. Caddemi, S. and Morassi, A. (2013). Multi-cracked Euler–Bernoulli beams: Mathematical modeling and exact solutions. International Journal of Solids and Structures, 50, 944–956.

140

Modern Trends in Structural and Solid Mechanics 1

Caddemi, S., Caliò, I., Cannizzaro, F. (2017). The dynamic stiffness matrix (DSM) of axially loaded multi-cracked frames. Mechanics Research Communications, 84, 90–97. Caliò, I., D’Urso, D., Greco, A. (2017). The influence of damage on the eigen-properties of Timoshenko spatial arches. Computers and Structures, 190, 13–24. Cannizzaro, F., Greco, A., Caddemi, S., Caliò, I. (2017). Closed form solutions of a multicarcked circular arch under static loads. International Journal of Solids and Structures, 121, 191–200. Capecchi, D., Ciambella, J., Pau, A., Vestroni, F. (2016). Damage identification in a parabolic arch by means of natural frequencies, modal shapes and curvatures. Meccanica, 51, 2847–2859. Cawley, P. and Adams, R. (1979). The location of defects in structures from measurements of natural frequencies. Journal of Strain Analysis, 14(2), 49–57. Cerri, M. and Ruta, G. (2004). Detection of localised damage in plane circular arches by frequency data. Journal of Sound and Vibration, 270, 39–59. Cerri, M., Dilena, M., Ruta, G. (2008). Vibration and damage detection in undamaged and cracked circular arches: Experimental and analytical results. Journal of Sound and Vibration, 314, 83–94. Chati, M., Rand, R., Mukherjee, S. (1997). Modal analysis of a cracked beam. Journal of Sound and Vibration, 207(2), 249–270. Ciambella, J. and Vestroni, F. (2015). The use of modal curvatures for damage localization in beam-type structures. Journal of Sound and Vibration, 340, 126–137. Dimarogonas, A.D. (1981). Buckling of rings and tubes with longitudinal cracks. Mechanics Research Communications, 8, 179–186. Dimarogonas, A.D. (1982). Crack identification in aircraft structures. Proceedings of the First National Aircraft Conference, pp. 11–22. Dimarogonas, A.D. (1996). Vibration of cracked structures: A state of the art review. Engineering Fracture Mechanics, 55(5), 831–857. Dimarogonas, A.D. and Paipetis, S. (1983). Analytical Methods in Rotor Dynamics. Applied Science, Paris. Doebling, S., Farrar, C., Prime, M., Shevitz, D. (1996). Damage identification and health monitoring of structural and mechanical systems from changes in their vibration characteristics: A literature review. Los Alamos Nat. Lab. Report, LA-13070-MS, pp. 1–129. Dugdale, D. (1960). Yielding of steel sheets containing slits. Journal of Mechanical Physics and Solids, 8, 100–104. Eroglu, U. and Ruta, G. (2020). Closed-form solutions for elastic tapered parabolic arches under uniform thermal gradients. Meccanica, 55, 1135–1152. Eroglu, U. and Tufekci, E. (2016). Exact solution based finite element formulation of cracked beams for crack detection. International Journal of Solids and Structures, 96, 240–253. Eroglu, U. and Tufekci, E. (2017a). Crack modeling and identification in curved beams using differential evolution. International Journal of Mechanical Sciences, 131–132, 435–450.

Buckling and Post-Buckling of Parabolic Arches with Local Damage

141

Eroglu, U. and Tufekci, E. (2017b). Free vibration of damaged frame structures considering the effects of axial extension, shear deformation and rotatory inertia: Exact solution. International Journal of Structural Stability and Dynamics, 17(10), 1750111. Eroglu, U., Ruta, G., Tufekci, E. (2019). Natural frequencies of parabolic arches with a single crack on opposite cross-section sides. Journal of Vibration and Control, 25(7), 1313–1325. Eroglu, U., Paolone, A., Ruta, G., Tufekci, E. (2020). Exact closed-form static solutions for parabolic arches with concentrated damage. Archive of Applied Mechanics, 90, 673–689. Gounaris, G. and Dimarogonas, A. (1988). A finite element of a cracked prismatic beam for structural analysis. Computers and Structures, 28(3), 309–313. Greco, A. and Pau, A. (2011). Detection of a concentrated damage in a parabolic arch by measured static displacements. Structural Engineering and Mechanics, 39(6), 751–765. Griffith, A. (1920). The phenomena of rapture and flow in solids. Philosophical Transactions of the Royal Society of London, Series A, 221(582–593), 163–198. Hubbard, J. and West, B. (1995). Differential Equations: A Dynamical Systems Approach. Springer-Verlag, New York. Inglis, C. (1913). Stresses in plates due to the presence of cracks and sharp corners. Transactions of the Institute of Naval Architects, 55, 219–241. Irwin, G. (1957a). Analysis of stresses and strains near the end of a crack traversing a plate. Journal of Applied Mechanics, 24, 361–364. Irwin, G. (1957b). Relation of stresses near a crack to the crack extension force. Proceedings of the 9th Congress of Applied Mechanics, Brussels, p. 245. Irwin, G. (1960a). Fracture Mechanics in Structural Mechanics. Pergamon Press, Oxford. Irwin, G. (1960b). Plastic zone near a crack and fracture toughness. Proceedings of the 7th Sagamore Ordnance Materials Conference, Syracuse University, New York. Ju, F. and Minovich, M. (1986). Modal frequency method in diagnosis of fracture damage in structures. Proceedings of the 4th International Modal Analysis Conference, Los Angeles, CA. Karaagac, C., Ozturk, H., Sabuncu, M. (2011). Crack effects on the in-plane static and dynamic stabilities of a curved beam with an edge crack. Journal of Sound and Vibration, 330(8), 1718–1736. Khiem, N. and Lien, T. (2001). A simplified method for natural frequency analysis of a multiple cracked beam. Journal of Sound and Vibration, 245(4), 737–751. Kirsch, E. (1898). Die theorie der elastizität und die bedürfnisse der festigkeitslehre. Zeitschrift des Vereines deutscher Ingenieure, 42, 797–807. Kirshmer, P. (1944). The effect of discontinuities on the natural frequency of beams. Proceedings of ASTM International, 44, 897–904. Kisa, M. and Arif Gurel, M. (2007). Free vibration analysis of uniform and stepped cracked beams with circular cross sections. International Journal of Engineering Science, 45(2–8), 364–380.

142

Modern Trends in Structural and Solid Mechanics 1

Krawczuk, M., Palacz, M., Ostachowicz, W. (2003). The dynamic analysis of a cracked Timoshenko beam by the spectral element method. Journal of Sound and Vibration, 264(5), 1139–1153. Mazanoglu, K., Yesilyurt, I., Sabuncu, M. (2009). Vibration analysis of multiple-cracked non-uniform beams. Journal of Sound and Vibration, 320(4–5), 977–989. Müller, W., Herrmann, G., Gao, H. (1993). A note on curved cracked beams. International Journal of Solids and Structures, 30, 1527–1532. Orhan, S. (2007). Analysis of free and forced vibration of a cracked cantilever beam. NDT & E International, 40(6), 443–450. Ostachowicz, W. and Krawwczuk, M. (1992). Coupled torsional and bending vibrations of a rotor with an open crack. Archive of Applied Mechanics, 62, 191–201. Pau, A., Greco, A., Vestroni, F. (2010). Numerical and experimental detection of concentrated damage in a parabolic arch by measured frequency variations. Journal of Vibration and Control, 17(4), 605–614. Peano, G. (1888). Intégration par séries des équations différentielles linéaires. Mathematische Annalen, 32(3), 450–456. Pease, M. (1965). Methods of Matrix Algebra. Academic Press, New York. Pestel, E. and Leckie, F. (1963). Matrix Methods in Elastomechanics. McGraw-Hill, New York. Pflüger, A. (1964). Stabilitätsprobleme der Elastostatik. Springer-Verlag, Berlin. Pignataro, M. and Ruta, G. (2003). Coupled instabilities in thin-walled beams: A qualitative approach. European Journal of Mechanics – A/Solids, 22(1), 139–149. Pignataro, M., Rizzi, N., Luongo, A. (1991). Stability, Bifurcation and Post-critical Behaviour of Elastic Structures. Elsevier, Amsterdam. Pignataro, M., Rizzi, N., Ruta, G. (2008). A beam model for the flexural-torsional buckling of thin-walled members. Thin Walled Structures, 46, 816–822. Rice, J. (1968). A path independent integral and the approximate analysis of strain concentration by notches and cracks. ASME Journal of Applied Mechanics, 35, 379–386. Rice, J. and Levy, N. (1972). The part-through surface crack in an elastic plate. Journal of Applied Mechanics, 39, 185–194. Simitses, G. and Hodges, D. (2006). Fundamentals of Structural Stability. ButterworthHeinemann, Burlington. Sinha, J.K. and Friswell, M.I. (2002). Simulation of the dynamic response of a cracked beam. Computers and Structures, 80(18–19), 1473–1476. Slavik, A. (2007). Product Integration, its History and Applications. Matfyz Press, Prague. Tada, H., Paris, P., Irwin, G. (2000). The Stress Analysis of Cracks Handbook, 3rd edition. ASME Press, New York. Thomson, W. (1949). Vibration of slender bars with discontinuities in stiffness. Journal of Applied Mechanics, 17, 203–207.

Buckling and Post-Buckling of Parabolic Arches with Local Damage

143

Timoshenko, S. and Gere, J. (1961). Theory of Elastic Stability. McGraw-Hill, New York. Timoshenko, S. and Goodier, J. (1951). Theory of Elasticity, 2nd edition. McGraw Hill, New York. Tufekci, E. and Arpaci, A. (2006). Analytical solutions of in-plane static problems for non-uniform curved beams including axial and shear deformations. Structural Engineering and Mechanics, 22(2), 131–150. Tufekci, E., Eroglu, U., Aya, S. (2017). A new two-noded curved beam finite element formulation based on exact solution. Engineering with Computers, 33(2), 261–273. Viola, E., Nobile, L., Federici, L. (2002). Formulation of cracked beam element for structural analysis. Journal of Engineering Mechanics, 128(2), 220–230. Wang, C., Wang, C., Reddy, J. (2005). Exact Solutions for Buckling of Structural Members. CRC Press, Boca Raton. Wells, A. (1963). Application of fracture mechanics at and beyond general yielding. British Welding Journal, 11, 563–570.

7 Inelastic Microbuckling of Composites by Wave-Buckling Analogy

7.1. Introduction Microbuckling in fiber-reinforced elastic composite materials, associated with the composite compressive strength, has been extensively studied over the years. A two-dimensional problem of a periodically layered bi-material, in which the stiff and soft layers represent the fibers and matrix, respectively, has been frequently used to model the composite. In the classical investigation by Rosen (1964), which has been summarized by Jones (1975), the two lower buckling modes were identified to be: the shear buckling mode characterized by in-phase deformation of fiber and matrix layers, and the transverse buckling mode with fiber and matrix layers exhibiting anti-phase deformation. Based on energy considerations, Rosen established an expression for the stress buckling in shear mode with very long wavelength (when compared to the thickness of the stiff fiber layer). Typically, in many studies of this problem, the stiff layer was modeled as an Euler–Bernoulli beam (see Waas et al. (1990), for example). This approach was also used by Parnes and Chiskis (2002), where a comprehensive list of references can be found. In this study, the infinite fiber layer was considered to be embedded in an elastic matrix, with the interactions between the layers deduced on the basis of the mechanics of materials. The results show that the long-wavelength shear buckling mode is adequate for composites of a relatively large fiber volume fraction, while composites of fiber volume fractions smaller than a specific transition value buckle in a finite wavelength shear mode. Guz (1969), Waas (1992), Nestorovic and Triantafyllidis (2004), Aboudi and Gilat (2006) and Gilat (2010), among others, used a full elasticity analysis of layered and fibrous composites, including the effects of combined loading as well as material anisotropy.

Chapter written by Rivka G ILAT and Jacob A BOUDI. Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

146

Modern Trends in Structural and Solid Mechanics 1

To better adjust theoretical predictions to experimentally measured composite compressive strength, models taking optional failure mechanisms, non-uniform pre-buckling stress distribution, fibers misalignment and waviness into account, as well as nonlinear and inelastic material behavior, were suggested by Budiansky and Fleck (1993), Kyriakides et al. (1995), Drapier et al. (1999), Guz et al. (2005) and Yokozeki et al. (2005), where further literature is referenced. In previous work, Aboudi and Gilat (2006) used the analogy between the governing equations for the analysis of microbuckling and the elastodynamic equations for wave propagation in solids, in order to predict the elastic microbuckling within a periodic layered bi-material and fiber-reinforced composites subjected to a uniaxial compressive load, parallel to the layer/fiber direction. In the present investigation, this wave-buckling analogy is used for the analysis of inelastic bifurcation microbuckling in layered composites. To this end, the instantaneous stiffness tensor of the inelastic layer is herein used in the bifurcation buckling analysis. The evolution of the instantaneous stiffness tensor of initially isotropic material under axial loading conditions, reveals that plasticity induces the softening and orthotropy of the material. In order to take this developing orientation dependency into account, the wave-buckling analogy is generalized for the case of orthotropic materials. It is applied to the rate version of the linearized buckling equation, in an incremental manner and in conjunction with constitutive relations reflecting the degradation of material stiffnesses. Assuming that the composite consists of elastic–viscoplastic materials, its behavior could have been conveniently described by the unified (Bodner and Partom 1975) model. However, while this viscoplastic constitutive law is well performing for proportional loading conditions, it cannot follow the plasticity induced decrease of shear modulus under uniaxial loading, and thus predicts this value as if it remains intact. As the importance of shear in the onset of instability in layered composites was indicated by the extensive work of the above-mentioned research and others, the modified constitutive law proposed by Rubin and Bodner (1995) for non-proportional buckling is adopted. This enables an adequate update of all the instantaneous stiffnesses ruling the onset of instability. 7.2. Buckling-wave propagation analogy Consider a periodic bi-material layered composite as shown in Figure 7.1(a). The repeating layered unit cell consists of two layers and is described with respect to global coordinates X1 , X3 , where X3 is the layering direction. In addition, a local (β) coordinate x3 , located at the bottom surface of each layer, β = 1, 2, is introduced. In a previous article by Aboudi and Gilat (2006), a wave-buckling analogy was used for the analysis of the microbuckling within such a periodic bi-material subjected to a uniaxial compressive load parallel to the layers direction (Figure 7.1(b)). According to this analogy, the buckling load can be directly extracted from the corresponding

Inelastic Microbuckling of Composites by Wave-Buckling Analogy

147

dispersion relation of the layered material, which provides the phase velocity of a wave in terms of the wavelength. 0(1)

X1

(2)

x3

X3 1 2 d1 d2

(a)

(b)

0(2)

Figure 7.1. (a) The repeating unit cell and (b) shear mode microbuckling of a periodic layered bi-material composite

The linear equations of motion governing the propagation of harmonic waves in a continuum are σjm,j = ρum,tt

[7.1]

j, m = 1, 2, 3

where σjm and um are the components of the stress tensor and the displacements vector, respectively, ρ is the mass density, t is the time, and the comma represents the differentiation. For the investigation of sagittal harmonic waves propagating in the layered composite, it is assumed that any field variable s(β) , in the β layer, has the form: (β)

s(β) = S (β) exp[ik(X1 + η (β) x3 − ct)]

[7.2]

where S (β) is an amplitude factor, k is the wave number associated with the (β) wavelength Λ = 2π/k, c is the phase velocity, x3 is the local layer coordinate in the (β) layering direction, η is an unknown parameter, and i is the imaginary unit. By using the form [7.2] in [7.1], it follows that the derivative with respect to X1 should be replaced by ik, the derivative with respect to x3 should be replaced by ikη and the derivative with respect to the time t is replaced by −ikc. Consequently, the following equations are established: (β)

(β)

(β)

(β)

(β)

(β)

ikσ11 + ikη (β) σ31 + ρ(β) (kc)2 u1 ikσ13 + ikη (β) σ33 + ρ(β) (kc)2 u3

=0 =0

[7.3]

148

Modern Trends in Structural and Solid Mechanics 1

On the contrary, the linearized equilibrium equations governing the bifurcation 0(β) buckling of a layer under pre-buckling uniaxial layer stress σ11 are (Whitney 1987) (β)

0(β) (β)

σjm,j − σjn um,nj = 0

[7.4]

j, m, n = 1, 3

For a classical cylindrical buckling, it is assumed that any field variable b(β) in the layer has the form (β)

b(β) = B (β) exp[ik(X1 + η (β) x3 )]

[7.5]

where B (β) is an amplitude factor and k is the buckling wave number which is associated with the wavelength Λ = 2π/k in the X1 direction. By substituting the form [7.5] in [7.4], the derivative with respect to X1 should be replaced by ik, and the derivative with respect to x3 should be replaced by ikη (β) . Accordingly, for the material occupying a layer β, [7.4] takes the form (β)

(β)

0(β)

(β)

(β)

(β)

0(β)

(β)

ikσ11 + ikη (β) σ31 + σ11 k 2 u1 ikσ13 + ikη (β) σ33 + σ11 k 2 u3

=0 [7.6]

=0

The comparison between [7.3] and [7.6] reveals that an analogy between wave propagation and buckling can be established and 0(β)

ρ(β) c2 ⇐⇒ σ11

[7.7]

Hence, the buckling load can be directly extracted from the corresponding dispersion relation of the layered material, which provides the phase velocity of a wave in terms of the wavelength. 7.3. Microbuckling in elastic orthotropic composites Aiming at the analysis of layered composites consisting of inelastic materials, we note that the properties of such materials develop orientation dependency upon loading. Focusing on buckling due to the overall uniaxial compression in the X1 direction, orthotropic constitutive relations should be taken into account for the layer material, with the material symmetry axes coinciding with the load direction. Hence, the stress–strain relations for the material occupying the β layer are ⎡

(β)

(β)

(β) σjn = Cjnpq pq ,

C(β)

C11 ⎢ C21 ⎢ ⎢ C31 =⎢ ⎢ 0 ⎢ ⎣ 0 0

C12 C22 C32 0 0 0

C13 C23 C33 0 0 0

0 0 0 C44 0 0

0 0 0 0 C55 0

⎤(β) 0 0 ⎥ ⎥ 0 ⎥ ⎥ 0 ⎥ ⎥ 0 ⎦ C66

[7.8]

Inelastic Microbuckling of Composites by Wave-Buckling Analogy

149

(β)

and the strains pq are related to the displacements through the following linear relations: 1 (β) (β) (β) pq = (up,q + uq,p ) [7.9] 2 To obtain the critical stress, the buckling curve is first obtained by following the approach suggested by Nayfeh (1995) for establishing the dispersion curves. The substitution of [7.8] and [7.9], in conjunction with the form [7.5] for the displacement components, into [7.6] reduces the latter to the following two algebraic equations: (β)

(β)

Knj Uj

= 0,

[7.10]

n, j = 1, 3

with (β)

(β)

(β)

0(β)

K11 = C11 + C55 (η (β) )2 + σ11 (β)

(β)

(β)

K13 = η (β) (C13 + C55 ) (β)

(β)

(β)

(β)

(β)

0(β)

K33 = C55 + C33 (η (β) )2 + σ11

[7.11]

Knj = Kjn (β)

(β)

where Uj are the displacement amplitudes. For a non-trivial solution, Uj , the determinant of K(β) must vanish, thus forming a characteristic equation of order four, defining four values of η (β) for each layer  (β) (β)2 (β) (β) −a ± a2 − 4a1 a3 2 (β) (β) (β) (β) (β) η1,3 = ±

, η2 = −η1 , η4 = −η3 (β) 2a1 (β)

a1

(β)

a2

(β)

(β)

= C33 C55 , (β)

(β)

(β)

a3

(β)

0(β)

(β)

0(β)

= (C55 + σ11 )(C11 + σ11 )

0(β)

(β)

(β)

0(β)

(β)

(β)

= C55 (C55 + σ11 ) + C33 (C11 + σ11 ) − (C13 + C55 )2 [7.12]

For each ηk , k = 1, 2, ...4, [7.10] can be used to define

(β)

(β) U3q K11 δq(β) = =− U1q K13

[7.13]

By using this definition in conjunction with the constitutive relations [7.8] and the kinematic relations [7.9], the displacements and stresses in a layer can be expressed as follows: ⎡ ⎤(β) ⎡ ⎤(β) u1 U11 ⎢ u3 ⎥ ⎢ ⎥ (β) (β) ⎢ U12 ⎥ ⎥ P(β) = ⎢ [7.14] ⎣ σ33 ⎦ = X D ⎣ U13 ⎦ exp[ikX1 ] σ31 U14

150

Modern Trends in Structural and Solid Mechanics 1

where X(β) and D(β) are 4 × 4 matrices with the entries (β)

(β)

X2q = δq(β)

X1q = 1, (β)

(β)

(β)

(β)

X3q = C13 + C33 ηq(β) δq(β) , (β) = 0, Dpq

(β)

X4q = C55 (ηq(β) + δq(β) )

p = q (β)

(β) = exp[ikηq(β) x3 ], Dpq

p = q,

[7.15]

p, q = 1, 2, ...4

The continuity constraints at the interface between adjacent layers of the representative unit and periodicity conditions can be cast to have the following form: (2)

P

(2)

(1)

(x3 =d2 )

= A(2) A(1) P

(1)

(1)

(x3 =0)

=P

[7.16]

(1)

(x3 =0)

with (β)

A(β) = X(β) D

(β)

(x3 =dβ )

(X(β) )−1

[7.17]

For a non-trivial solution, it is required that det[A(2) A(1) − I] = 0

[7.18] 0(β)

Recalling that A(β) = A(β) (k, σ11 ) and that under the uniform strain loading condition 0(β)

σ11

0 = σ11

(β) C¯11 (d1 + d2 ) , (1) (2) C¯ d1 + C¯ d2 11

11

(β)

(β)

(β)

(β)

C C −C C (β) C¯11 = 11 33 (β) 13 13 C33

[7.19]

0 [7.18] constitutes the relations between k and σ11 . For each wave number k, the 0 smallest root of [7.18] is σ ˜11 , which is the buckling stress. The corresponding (β) eigenvector Uj represents the buckling mode of wavelength Λ = 2π/k. The results of numerous previous works, showing that this buckling mode is the shear mode shown in Figure 7.1(b), make our choice of the considered representative unit 0 legitimate. Microbuckling occurs under the minimal value σ11cr , with wavelength 0 Λcr , minimizing σ ˜11 (Λ).

7.4. Inelastic microbuckling To consider inelastic bifurcation microbuckling in layered composites, the layer material is assumed to be an elastic–viscoplastic material and a constitutive law, suitable for the description of its behavior, is adopted. To explain the consideration guiding our choice of a material constitutive law, we first mention that simple expressions for the microbuckling stress in layered materials

Inelastic Microbuckling of Composites by Wave-Buckling Analogy

151

were proposed in the pioneering and numerous subsequent works dealing with the subject. Relating to a bi-material composite with a repeating unit cell consisting of one layer occupied by a stiffer material and termed fiber (subscript f ) and a second softer layer termed matrix (subscript m), the long-wavelength limit from Rosen (1964) for elastic materials is 0 σ11cr =

μm 1 − vf

[7.20]

where μm is the matrix shear modulus and vf is the volume fraction of the fiber material. One of the modifications suggested by Budiansky and Fleck (1993) for composites including a plastic matrix layer is given by

μIm Ems Em μm 0 σ11cr = = [7.21] 1 − vf Ems Em + 3μm (Em − Ems ) 1 − vf where Em and Ems are the matrix elastic and the secant moduli, respectively, and the superscript I stands for inelastic. Both the above expressions testify the significant role of the matrix shear modulus in the microbuckling of layered composites. Furthermore, inelastic material behavior reveals softening, which can be viewed as the reduction of the material mechanical properties. Hence, it is concluded that the effect of plasticity on the material shear modulus might be of importance when analyzing the inelastic microbuckling behavior of composites. In order for this effect to be reflected, the model proposed by Rubin and Bodner (1995) for non-proportional buckling is used. This model, which is based on the assumption that the plastic strain is explicitly related to the total strain rate, is a modification of the Bodner and Partom (1975) elastic–viscoplastic model. While in the framework of the latter, the shear modulus is not affected by plasticity evolved under axial loading, the modified model is capable of predicting the plasticity induced reduction of the shear modulus due to such a loading. The adopted constitutive law is 

VP σ˙ ij = (LE ijkl + Lijkl )˙kl − Γσij

[7.22]



where ˙ij is the total strain rate, σij is the deviatoric stress, and Γ is a scalar function VP involved in the flow rule. LE ijkl is the elastic part of the stiffness tensor and Lijkl is its viscoplastic part, which enables the detection of changes in the shear stiffness in the absence of direct shear stress and which is unique for the present model. This constitutive relation can be written in the following common form: I σ˙ ij = Cijkl ˙kl

[7.23]

152

Modern Trends in Structural and Solid Mechanics 1

I such that Cijkl represents the instantaneous stiffness of the elastic–viscoplastic material. In order to determine the inelastic microbuckling, we use the rate version of the buckling equations [7.4] (β)

0(β) (β)

σ˙ jm,j − σjn u˙ m,nj = 0

j, m, n = 1, 3,

β = 1, 2

[7.24]

Following the procedure for elastic buckling, the inelastic buckling stress is obtained in an incremental manner as follows. At each loading increment, namely for 0 each incrementally increased value of σ11 , the instantaneous stiffnesses of the inelastic phase are computed. For each value of k, the instantaneous matrix K(β)I , [7.11] is constructed, and the instantaneous characteristic equation det[K(β)I ] = 0 is (β) solved to determine the values of ηq , β = 1, 2, q = 1, ...4. The instantaneous matrices A(β)I are then built and the instantaneous buckling condition [7.18] det[A(2)I A(1)I − I] = 0

[7.25]

is checked. The smallest stress for which the instantaneous determinant vanishes is 0 the buckling stress σ ˜11 with a mode of wavelength Λ = 2π/k. The critical inelastic bifurcation microbuckling stress is the minimum of the latter values over k. 7.5. Results and discussion To demonstrate the validity of the present approach, results for the elastic case are first examined. In the following, the volume fraction vf defines the portion occupied by the material of layer 1. The effect of the stiffness ratio E (1) /E (2) (where E (β) is the β layer material Young’s modulus) on the critical buckling strain is shown in Figure 7.2. The present results are close to those extracted from the works of Waas (1992) and Parnes and Chiskis (2002). Further results assessing the validity of the wave-buckling analogy and the adoption of Nayfeh’s wave propagation analysis can be found in Aboudi and Gilat (2006) and Gilat (2010). The inelastic microbuckling in a composite consisting of elastic stiff layers and elasto-viscoplastic soft layers is presented in Figures 7.3–7.5. Specifically, the B/Al composite is considered with elastic shear moduli μ(1) = 153.8GP a and μ(2) = 26.6GP a and Poisson’s ratios ν (1) = 0.3 and ν (2) = 0.335, for boron and aluminum, respectively. The inelastic material parameters, characterizing the inelastic aluminum layer (see Rubin and Bodner (1995)), are Γ0 = 108 s−1 , κ0 = 0.4GP a, n = 1., m1 = 1000GP a−1, m2 = 40000GP a−1, Z1 = 0.53GP a, Z3 = 0.3GP a, b = 10−6 s−1 , f1 = 0.23 and f2 = 0.7.

Inelastic Microbuckling of Composites by Wave-Buckling Analogy

153

Figure 7.2. Effect of layer stiffness ratio on the elastic critical bifurcation microbuckling strain, vf = 0.05. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Figure 7.3. Buckling curves exhibiting the volume fraction-dependent variation of microbuckling stress with the buckling wavelength for B/Al resulting from (a) elastic (b) inelastic analyses. For a color version of this figure, see www.iste.co.uk/challamel/ mechanics1.zip

Figure 7.3 presents the buckling curves corresponding to the dispersion curves of the analogical wave propagation problem. To reveal the effect of plasticity, curves resulting from elastic and inelastic analyses are given for various values of volume fractions. For each buckling wavelength, these curves show the corresponding smallest buckling stress (smallest root of [7.25]). For each volume fraction, the

154

Modern Trends in Structural and Solid Mechanics 1

minimum point of the curve defines the critical bifurcation microbuckling stress and wavelength. As expected, for both elastic and elasto-plastic cases, composites of a small volume fraction undergo small wavelength microbuckling, while those of a large volume fraction yield long-wavelength buckling. Plasticity is observed to decrease the volume fraction at which the transition occurs from short-wave buckling to long-wave buckling. It also dramatically reduces the critical microbuckling stresses. The dependence of the critical buckling stress and wavelength on the composite volume fraction is presented in Figure 7.4. In addition to the present elastic results (dashed line), the long-wave approximation of Rosen (1964) [7.20] is also shown. The present inelastic buckling stresses (solid line) are somewhat higher than those predicted by [7.21], Budiansky and Fleck (1993). A similar trend can be deduced from the results presented by Yokozeki et al. (2005) for the special case of almost aligned linear elastic fibers embedded in an elastic–plastic matrix. Both present elastic and inelastic curves exhibit a tendency change at the point of transition from finite to infinite buckling wavelength.

Figure 7.4. Effect of the volume fraction on (a) the critical microbuckling stress and (b) the critical buckling wavelength of B/Al. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

The presence of this transition point can also be identified by the solid curve presented in Figure 7.5(a), showing the variation of the critical microbuckling strain with the volume fraction, as predicted by the presently suggested approach. This is in contrast with the qualitative behavior given by [7.21] which, being based on Rosen’s expression, cannot reflect the essence of finite wavelength buckling in composites of a low volume fraction.

Inelastic Microbuckling of Composites by Wave-Buckling Analogy

155

Figure 7.5. (a) Effect of the volume fraction on the critical microbuckling strain in the B/Al composite and (b) plasticity-induced reduction of the Al shear modulus. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1. zip

The plasticity-induced change of the inelastic phase shear modulus, defined by the presently adopted constitutive law, is indicated by the solid line shown in Figure 7.5(b). This shows a significant reduction at the level of critical buckling strain. As observed with respect to the buckling stress, the values of the presently predicted modified shear modulus are less conservative than those defined by [7.21]. To conclude, we examine the behavior of a composite in which the stiffer layer is inelastic while the softer is assumed to be linearly elastic. Specifically, Al/epoxy is considered with μ(1) = 26.6GP a, μ(2) = 1.96GP a, ν (1) = 0.335 and ν (2) = 0.328 and with the inelastic material properties as given above. The behavior of such composites is presented in Figure 7.6. Figure 7.6(a) reveals that in contrast to the qualitative similarity between the elastic and inelastic behaviors exhibited in Figure 7.3, in the present case, there is a qualitative difference between the elastic and inelastic behaviors. According to Figure 7.6, inelasticity of the stiffer layer eliminates the short-wavelength buckling to long-wavelength buckling transition. This can be attributed to the plasticity-induced decrease of the inelastic material stiffness, decreasing the layer stiffness ratio E (1) /E (2) . As shown by Parnes and Chiskis (2002), for smaller stiffness ratios, the short- to long-wavelength transition occurs at a larger volume fraction. As indicated by Figure 7.2, such a reduction of the layers stiffness ratio yields buckling under larger critical buckling strains. It should be noted that under large strains, the assumption of linearly elastic behavior of the softer layer is questionable.

156

Modern Trends in Structural and Solid Mechanics 1

Figure 7.6. Effect of the volume fraction on (a) the critical microbuckling stress and (b) the critical buckling wavelength of Al/epoxy. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

In summary, in the present work, the inelastic bifurcation microbuckling in a bi-material periodic layered composite was formulated as a micromechanical problem. A solution was obtained by using the analogy between the buckling and wave propagation problems, in conjunction with an incremental approach, while adopting an elastic–viscoplastic constitutive law indicating plasticity reduction of all the involved moduli. 7.6. References Aboudi, J. and Gilat, R. (2006). Buckling analysis of fibers in composite materials by wave propagation analogy. Int. J. Solids Struct., 43, 5168–5181. Bodner, S.R. and Partom, Y. (1975). Constitutive equations for elastic-viscoplastic strain-hardening materials. ASME J. Appl. Mech., 42, 385–389. Budiansky, B. and Fleck, N.A. (1993). Compressive failure of fibre composites. J. Mech. Phys. Solids, 41, 183–211. Drapier, S., Grandidier, J., Poiter-Ferry, M. (1999). Towards a numerical model of the compressive strength for long fibre composites. Eur. J. Mech. A/Solids, 18, 69–92. Gilat, R. (2010). A 3D thermoelastic analysis of the buckling of a layer bonded to a compliant substrate and related problems. Int. J. Solids Struct., 47, 2533–2543. Guz, A.N. (1969). On setting up a stability theory of unidirectional fibrous materias. Institute of Mechanics, Academy of Sciences of the Ukrainian SSR, Prikladnaya Mekhanika, 5, 62–70. Guz, A.N., Rodger, A., Guz, I. (2005). Developing a compressive failure theory for nanocomposites. Int. Appl. Mech., 41, 233–255. Jones, R. (1975). Mechanics of Composite Materials. Scripta Book Company, Washington, DC.

Inelastic Microbuckling of Composites by Wave-Buckling Analogy

157

Kyriakides, S., Arseculeratne, R., Perry, E.J., Liechti, K.M. (1995). On the compressive failure of fiber reinforced composites. Int. J. Solids Struct., 32, 689–738. Nayfeh, A. (1995). Wave Propagation in Layered Anisotropic Media with Applications to Composites. Elsevier, Amsterdam. Nestorovic, M. and Triantafyllidis, N. (2004). Onset of failure in finitely strained layered composites subjected to combined normal and shear loading. J. Mech. Phys. Solids, 52, 941–974. Parnes, R. and Chiskis, A. (2002). Buckling of nano-fiber reinforced composites: re-examination of elastic buckling. J. Mech. Phys. Solids, 50, 855–879.

A

Rosen, B. (1964). Mechanisms of Composite Strengthening. In Fiber Composite Materials. American Society of Metals, Cleveland, OH. Rubin, M.B. and Bodner, S.R. (1995). An incremental elastic–viscoplastic theory indicating a reduced modulus for non-proportional buckling. Int. J. Solids Struct., 32, 2967–1987. Waas, A.M. (1992). Effect of interphase on compressive strength of unidirectional composites. J. Appl. Mech., 59, s183–s188. Waas, A.M., Babcock, C.D., Knauss, W.G. (1990). A mechanical model for elastic fiber microbuckling. ASME J. Appl. Mech., 57, 138–149. Whitney, J. (1987). Structural Analysis of Laminated Anisotropic Plates. Pergamon Press, Oxford. Yokozeki, T., Ogasawara, T., Ishikawa, T. (2005). Effects of fiber nonlinear properties on the compressive strength prediction of unidirectional carbon–fiber composites. Compos. Sci. Technol., 65, 2140–2147.

8 Quasi-Bifurcation of Discrete Systems with Unstable Post-Critical Behavior under Impulsive Loads

The huge economic and environmental consequences due to explosions in the oil industry have recently motivated the study of thin-walled cylindrical shells subjected to impulsive waves due to an explosion. One of the main challenges in this field is the identification of suitable dynamic buckling criteria for this class of loads. This paper focuses on a simple geometrically nonlinear two degree-of-freedom system subjected to impulsive loads of very short duration with respect to their fundamental periods. Similar to what occurs in shell structures, the model includes both membrane and bending stiffness components and displays an unstable static behavior at the critical state. The description of the motion is formulated using Lagrange equations and they are integrated numerically by means of an implicit code. By analogy with the static case, the dynamic path that follows the motion of a degree of freedom at a given load level is defined as the fundamental motion; any perturbed path that departs from the fundamental path is interpreted as a bifurcation of the original motion and termed quasi-bifurcation by Lee. Following the original criterion due to Lee, a coefficient calculated as the projection of the perturbed motion on the corresponding perturbed acceleration is investigated, and instability is said to occur for positive values of the coefficient. The present results show that the quasi-bifurcation criterion, as originally presented by Lee, is a necessary condition but may yield low values of instability loads because a positive projection may occur for a short time before the structure returns to a stable condition with a negative coefficient. A modification is proposed in which the time integration of the inner product due to Lee is carried out, and the analysis shows that this allows a

Chapter written by Mariano P. AMEIJEIRAS and Luis A. GODOY. Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

160

Modern Trends in Structural and Solid Mechanics 1

clearer identification of bounds in the response that are associated with dynamic buckling. 8.1. Introduction Explosions have been identified as the most frequent accidents affecting the oil industry (Kang et al. 2016). Interest in the behavior of oil storage tanks under blast loads caused by explosions has intensified over the last decade following the notorious accidents at a tank farm located in Buncefield, UK, in 2005 and another one in Bayamon, Puerto Rico, in 2009. Understanding and modeling this class of problems requires computation of the nonlinear dynamic response under short duration impulsive loads. Although such analysis is part of the state of the art in analytical and finite element approaches (see, for example, Nayfeh and Mook (1979); Nayfeh and Balachandran (2004)), such studies may sometimes be obscured by the complexity of the problem under consideration, whereas the interpretation of results and the evaluation of stability to identify dynamic buckling are topics still subject to discussion. Thus, to improve our understanding of the mechanics of behavior of shells under blast loads, it may be convenient to perform studies on much simpler analogous structural systems, for which the response can be modeled using analytical tools. The buckling of structures under step or impulsive loading has been the subject of research since the 1960s. Reviews of dynamic buckling criteria were presented, for example, in the books by Simitses (1990), Amabili (2008) and Kubiak (2013). This area of research falls within the field of non-classical problems in the theory of stability, as defined by Elishakoff et al. (2001). By analogy with the static case, we will define that a fundamental motion is the trajectory in a plot of a representative degree of freedom (DOF) with respect to time, computed at a given load level. Two forms of dynamic buckling, which are relevant to the present study, have been identified in the literature: (a) Instability in the same mode of the fundamental motion, associated with large amplitude oscillations. Forced vibrations are investigated in this approach using nonlinear dynamics, and a criterion is employed to assess the occurrence of unstable oscillations. The most common criterion employed in the literature was originally proposed by Budiansky and Roth (1962), in which the dynamic response is first computed for various load levels and the load versus amplitude of a representative DOF is plotted. By analogy with the static case, this plot is here called a pseudo-equilibrium path. Dynamic buckling is identified whenever there is a jump in the pseudo-equilibrium path, i.e. when a small increment in load causes a

Quasi-Bifurcation of Discrete Systems

161

non-proportional (large) increment in transient displacements. This criterion was originally developed for step loads, but its application to impulsive loads (such as those associated with explosions) is less evident. (b) Instability in a mode which is orthogonal to the fundamental motion. This behavior is known as quasi-bifurcation from a fundamental trajectory and was originally identified by Lee (1977, 1981). As stated before, the motion of an undisturbed system shows oscillations in a mode here identified as the fundamental mode. But there are cases in which a second mode grows at a critical time, leading to a quasi-bifurcation; at first, this takes the form of small random oscillations, but a divergent motion is finally achieved. Lee (1981) also identified that the dynamic post-bifurcation mode may change in time. The identification of a quasi-bifurcation state may be carried out either: (a) by establishing conditions at which dynamic bifurcation should occur, such as in the works of Kounadis et al. (1989) and Simitses (1990), or (b) by following stability indicators along the fundamental motion, such as frequencies of vibration, static eigenvalue problems, tangent matrix or static stability coefficients. Various authors in Europe (notably Kroplin and Dinkler 1986; Kleiber et al. 1987; Burmeister and Ramm 1990; Kratzig and Eller 1992) implemented strategies to evaluate stability along the fundamental motion. The dynamic buckling of a simple, two DOF model is investigated in this work under an impulsive load in order to explore the type of analysis that should be considered when addressing shell problems under blast pressures due to explosions. The framework of the present analysis is that given by Lee, and shortcomings are highlighted together with ways to improve the identification of stability of a motion. The chapter focuses on quasi-bifurcation phenomena, which is also believed to occur in shells under impulsive loads due to an explosion and which may be instrumental in explaining the failure modes observed in oil infrastructure. 8.2. Case study of a two DOF system with unstable static behavior In order to investigate the dynamic instability criterion developed by Lee (1977), a simple nonlinear geometric and elastic two DOF system is considered in this paper, for which a semi-analytical formulation may be implemented. The specific mechanical system was originally studied by Croll and Walker (1972) while addressing static stability. Interest in this model is twofold: first, it contains bending and membrane mechanisms to equilibrate the load, in much the same way as in shell structures; and second, its static behavior is characterized by an unstable critical state with imperfection sensitivity, as shown by Croll and Walker (1972).

162

Modern Trends in Structural and Solid Mechanics 1

The three rigid bars shown in Figure 8.1 have length L and their initial position is given by α0 and β0. Bending stiffness is provided by rotational springs Kf, whereas membrane stiffness is controlled by the extensional spring Km. The mass of the system is lumped at the joints with value M. The generalized DOF of the problem are rotations (t) and (t), also shown in Figure 8.1.

Figure 8.1. Two DOF model investigated in this work

The kinetic energy T, elastic potential V, dissipation D and generalized forces Q (excluding viscous damping) are given by:

1 M (U112 + U12 2 + U 212 + U 22 2 ) 2 1 V = ( K mU 312 + K f Δγ 12 + K f Δγ 2 2 ) 2 1 D = C (U12 2 + U 212 ) 2 ∂U 22  P  ∂U Q1nc =  12 + ∂α  2  ∂α T=

Q2nc =

[8.1]

P  ∂U12 ∂U 22  +   2  ∂β ∂β 

with displacement components given by: U11 = L  − cos ( β 0 − β 0ξ ) + cos ( β 0 − β 0ξ − α ) 

[8.2a]

U12 = L sin ( β 0 − β 0ξ ) − sin ( β 0 − β 0ξ − α ) 

[8.2b]

Quasi-Bifurcation of Discrete Systems

163

U 21 = − L cos ( β 0 − β 0ξ ) + L cos ( β 0 − β 0ξ − α ) − 1 2

{ }+ L {1 −  − sin ( β − β ξ − α ) + sin ( β − β )  } L 1 − sin ( β 0 ) − sin ( β 0 − β 0ξ ) 

2

2

0

0

[8.2c] 1 2

0

U 22 = L sin ( β 0 ) − sin ( β 0 − β ) 

[8.2d]

The scalar ξ denotes the amplitude of geometric imperfection, and is given by:

ξ=

α 0 − β0 β0

[8.3]

The angle changes at the hinges between bars are given by:     sin ( β 0 ) − sin ( β 0 − β 0ξ ) + Δγ 1 = − arctan  1  2 2   1 − sin ( β 0 ) − sin ( β 0 − β 0ξ )       − L sin ( β 0 − β 0ξ − α ) + L sin ( β 0 − β ) + arctan   2  L 1 −  − sin ( β 0 − β 0ξ − α ) + sin ( β 0 − β )  

}

   +α 1  2  

    sin ( β 0 ) − sin ( β 0 − β 0ξ ) − Δγ 2 = arctan  1  2 2   1 − sin ( β 0 ) − sin ( β 0 − β 0ξ )       − L sin ( β 0 − β 0ξ − α ) + L sin ( β 0 − β ) − arctan   2  L 1 −  − sin ( β 0 − β 0ξ − α ) + sin ( β 0 − β )  

}

  +β 1  2  

{

}

{

{

{

}

[8.4a]

[8.4b]

164

Modern Trends in Structural and Solid Mechanics 1

The equations of motion for this system are derived by use of Lagrange’s formulation (see, for example, Baruh 1999): ____ d ∂T ∂T ∂D ∂V − + + = Qinc , i = 1, n dt ∂qi ∂qi ∂qi ∂qi

[8.5]

in which qi are the generalized coordinates. For quasi-static systems, equation [8.5] takes the form of the principle of minimum potential energy and might be used to define the static equilibrium path including geometric nonlinearities. The resulting dynamic equations are cumbersome and for reasons of brevity have not been included in this work. The full version of the analytical formulation may be found in Ameijeiras (2020). The equations of motion are integrated numerically through the IDA code in the SUNDIALS package (Hindmarsh et al. 2005) within the Mathematica software (Wolfram Mathematica 2019). 8.3. Exploring the static and dynamic behavior of the two DOF system

The static and dynamic response of the two DOF model is explored in this section by assuming the data adopted by Croll and Walker (1972): 0=0.2 rad, 0=0.2 rad, L0=1 m, Kf=1 MN/m, Km=1000 MN/m, M=10 MNs2/m, and no viscous damping has been assumed (C=0) in the results presented in this chapter. The load P(t) is given by a triangular impulse:   t   P0  1 −  for 0 ≤ t ≤ t0 P (t ) =   t0   0 Otherwise 

[8.6]

where t0 is the duration of the positive phase of the impulsive load (assumed in this case as 0.05 s) and P0 is the peak force. As reference values, the natural periods of vibration of the system are computed as T1 = 6.71 s, T2 = 2.22 s, so the duration of the impulse is very small in comparison with the period of both modes, i.e. t0/T1=0.0075 and t0/T2=0.0225. Considering the static response, the case with ξ=0 displays a linear bifurcation at PB=1.18 MN, B=B=0.00764 rad, and the energy stored at this bifurcation state is V=0.0046 MNm. The geometrically nonlinear static response of the system with imperfections has been computed using Riks’ algorithm and the equilibrium paths are shown in Figure 8.2(a), where for increasing values of imperfection amplitude ξ the equilibrium paths show a maximum in each curve. As stated before, interest in this system is associated with its unstable static behavior at the critical state, which

Quasi-Bifurcation of Discrete Systems

165

is what occurs in some shell structures. The imperfection-sensitivity plot is shown in Figure 8.2(b) for the values of maximum in each curve normalized with respect to the bifurcation load, and this illustrates a behavior that is typical of some shell structures.

a)

b)

Figure 8.2. Geometrically nonlinear static response of the two DOF system shown in Figure 8.1. (a) Equilibrium paths for perfect and imperfect configurations. (b) Imperfectionsensitivity curve. For a color version of this figure, see www.iste.co.uk/challamel/ mechanics1.zip

Information about the potential dynamic behavior may be obtained by exploration of the strain energy at the unloaded state P=0, as it would occur in a system subjected to an impulse (Simitses 1990). The potential energy surface for P=0 is shown in Figure 8.3(a) in a three-dimensional view and in Figure 8.3(b) by means of contour curves. The plots show a minimum in the energy at points O and D, at which stable behavior occurs, and equilibrium states at A, B and C exhibiting a maximum (unstable behavior). Point O occurs at ==0, with V0=0, whereas point D occurs at ==0.4 rad, with V0=0.16 MNm, which is an energy level much higher than that required to reach PB. A local maximum occurs at =0.2 rad, =0.2 rad (V0=0.84 MNm) for point A; =0.4 rad and =0.2 rad (V0=0.18 MNm) for point B; and =0.2 rad, =0.4 rad (V0=0.18 MNm) for point C. For sudden loads and considering the impulse–moment theorem, it is possible to associate load levels with configuration points with known energy. Thus, providing an energy input V0=0.84 MNm in the form of an impulse should be sufficient to move the system from O to A causing an escape of the

166

Modern Trends in Structural and Solid Mechanics 1

oscillation in the vicinity of the origin. This case was identified by Simitses (1990) as a minimum guaranteed critical load (MGCL) and the associated load in this case is PMGCL=231.9 MN. Lower energy levels are required to move the system from O to B (or from O to C); such states were identified by Simitses as minimum possible critical load (MPCL) and in this case they occur at PMPCL=107.3 MN. Such load values are much larger than those found for the static PB.

a)

b)

Figure 8.3. Potential energy under P=0. (a) 3D view. (b) Iso lines. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Next, the dynamic response under an impulsive pressure load is explored. The response of the perfect system (ξ=0) for a load level PPMPCL assuming ξ=0. (a) Transit response. (b) Phase plane for α. (c) Phase plane for β. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

The visualization through the phase plane is a useful tool for a few DOF; however, generalization to multiple DOF becomes problematic. Not just the early transient response needs to be computed, sometimes it requires the computation of a few hundred seconds in order to visualize reaching a new attractor.

8.4. The dynamic stability criterion due to Lee Lawrence H.N. Lee (1923–2007), who was a professor at the University of Notre Dame in Indiana, USA, postulated in 1977 that if a perturbation ζ is imposed onto a system, the dynamic response reaches an unstable configuration whenever the projection of the perturbed acceleration vector on the perturbed displacement vector is positive, i.e. both acceleration and displacement are related in a positive way (Lee 1977). However, the configuration is stable if the projection is negative. Thus, if a positive change in displacement causes a positive change in acceleration, then the system will increase the amplitude of the motion.

168

Modern Trends in Structural and Solid Mechanics 1

According to Lee (1981, p. 81, equation 15), assuming a perturbed motion given by (p=ζ+δζ), the following condition applies for instability:

CL = δζr ( t ) δζ r ( t ) > 0 (sum in r ), t > tcr

[8.7]

where summation extends to the number of DOF of the system. The result in equation [8.7] is a scalar quantity, which is here called Lee’s coefficient CL. According to Lee (1981, p. 81), when CL>0 after a certain time tcr, for any possible deviated motion, the response  grows without bounds and the known motion is unstable. In this way, the system is stable up to time tcr or, in other words, the fundamental motion is bounded or stable when the perturbed force field remains a restoring or retarding one. The variables (t) and (t) describe the motion from the initial position 0 and 0, i.e. (0)=0 and (0)=0, and p and p designate the motion with a perturbed initial condition (0)= and (0)=. This reference can be associated with equation [8.7] as follows:

α = ζ 1 , α p − α = δα = δζ 1 β = ζ 2 , β p − β = δβ = δζ 2

[8.8]

To illustrate the use of the criterion proposed by Lee, consider the two DOF system under study, at a load level P0=110 MN>PMPCL. The transient response of a perfect system without damping is shown in Figure 8.5(a) considering both unperturbed (α, β) and perturbed (αp, βp) motions. The two DOF initially oscillate together but at t=3 s the mode changes and each DOF oscillates following a different pattern. The product CL due to Lee, shown in Figure 8.5(b), is initially negative (stable) and becomes positive at t≈1 s and again at t≈3 s, thus indicating an instability. It may be seen in Figure 8.5 that the two DOF depart in their motion in the perturbed system. The conclusion is that Lee’s criterion was capable of identifying that the system is unstable at this load level and that it has already gone through a quasi-bifurcation. Next, consider a lower load level, say P0=70 MN 0 (where Im( ) denotes the imaginary part of ). However, as damping is absent in the present mathematical model, all α ≡ 0 ( and so Im( ) ≡ 0) for load , and the vibrations are only neutrally stable. Flutter is values below the critical, initiated when two eigensolutions (eigenvalues and eigenvectors) coalesce. Instability is initiated already at the coalescence point because the differential equation [10.2] here has a solution in the form ( , ) =

( )+

( ) exp(i

),

[10.11]

where ( ) is the so-called associated eigenfunction. After this point, i.e. at a slightly higher load value than the critical one, the before double eigenvalue splits

Shape-optimized Cantilevered Columns

into two complex ones in the form grow exponentially in time. Let

,

209

= ± + i , ≠ 0. The oscillations thus

be the eigenvector of the adjoint system = .

[10.12]

The flutter criterion can then be written (Plaut 1972; Langthjem and Sugiyama 2000) in the form = 0, where the index +

[10.13]

indicates the onset of flutter. This also implies that = 0.

[10.14]

In the numerical computations, the flutter points were determined by using a bisection method. The criterion [10.13] was used as a check. 10.3.2. Optimum design The intended aim of the experiments was to verify both (I) critical load maximization, keeping the volume constant and (II) volume minimization, keeping the critical load constant. Without end mass and non-zero geometrical constraints, these two formulations are equivalent, meaning they give solutions which are identical, except for a scaling factor. This is not so for the present problem. 10.3.2.1. Optimization problem I: maximization of critical load under a constant volume , ,⋯, be the “flutter loads” at which pairs of eigensolutions Let coincide, and let be the smallest of these values. For the uniform column, it is . Figure 10.5 defines the multiple critical loads assumed that < < ⋯ < , , ⋯. In the case shown, the critical load equals the minimum critical load . Optimization problem I is to determine the design vector follows: Maximize critical load

=

( )

and can be posed as

210

Modern Trends in Structural and Solid Mechanics 1

subject to i)



, ,⋯,

ii)





iii) Volume = iv)



>

,

, = 1, 2, ⋯ ,

+ 1,

= constant, 0 for = 1, 3, ⋯ , 2 for = 2, 4, ⋯ , 2

− 1, for all

= 1, 2, ⋯ ,

+ 1,

0 for = 1, 3, ⋯ , 2 for = 2,4, ⋯ , 2

−1 for all < − 2

.

[10.17]

The inclusion of the constraint on the load–frequency curves, [10.15 (iv)] and [10.17 (iii)], is necessary to ensure a robust optimization algorithm which is able to provide good optimization design for any set of system parameters (Langthjem and Sugiyama 1999b, 2000b). However, in the present case, where the mass ratio = (mass of rocket motor)⁄(mass of column) ≈ 4.6, this constraint did not become active (see also section 10.3.2.4). 10.3.2.3. Solution by sequential linear optimization The nonlinear optimization problems [10.15] and [10.17] are linearized and solved iteratively by using sequential linear optimization (linear programming). . The ( + 1)th optimal After iterations, the design is specified by vector = + ∆ , is written redesign problem, for determining the design vector as follows: Maximize critical load

=

( )

subject to i)





+ ∑



ii)



+ ∆



, = 1, 2, ⋯ ,

iii)



iv)

= , −







, = 1, 2, ⋯ , + 1,

,

212

Modern Trends in Structural and Solid Mechanics 1



+ ∑



0 for = 1, 3, ⋯ , for = 2, 4, ⋯ ,

>

for ∗

=

, ∗ , ⋯ ,



,

[10.18]

for problem [10.15], and ) = min(

Minimize volume = min(

+ ∆

) = min( ∆ )

subject to i) ii)





+ ∑



+ ∆



iii)





≥ ,



+ ∑



, = 1, 2, ⋯ ,

+ 1 ,

− 0 for = 1, 3, ⋯ , for = 2, 4, ⋯ ,

>

for =



, ∗ , ⋯ ,



,

[10.19]

for problem [10.17]. The discrete load values = ∗ , ∗ , ⋯ , ∗ in the constraints [10.18 (iv)] and [10.19 (iii)] are in the range 0 < < . The ⁄ , and the derivative of the flutter loads with respect to the design variables, ⁄ derivative of the frequencies, , are given by :

= − and

=



[10.20]

respectively. The size of the design change is governed by move limits ∆ ∆ , = 1, 2, ⋯ , + 1, such that 0 ≥ ∆

≤ ∆

≤ ∆

≥ 0.

,

[10.21]

Shape-optimized Cantilevered Columns

213

The linearized problems [10.18] and [10.19] are converted into a pure linear programming problem by introduction of the positive variables ∆

= ∆

− ∆

≥ 0.

[10.22]

In terms of these variables, the design change limits are given by 0 ≤ ∆

≤ ∆

− ∆

.

[10.23]

The sequence of linear programming problems [10.18] and [10.19] are solved, for = 1, 2, ⋯, by applying the Simplex method (Press et al. 1992). During the iterations, the move limits are decreased automatically according to the size of the gradients (Press et al. 1992). Further computational details are given in Langthjem and Sugiyama (2000b, 2000c). 10.3.2.4. Optimality criterion by a single (simple) flutter load It was found in Langthjem and Sugiyama (1999b) that for columns equipped with a rocket motor of a certain size, such that the mass ratio ≳ 1, convergence of the optimum design is reached by a single (simple) flutter load. It was shown that the optimum design then satisfies the optimality criterion (Pedersen and Seyranian 1983) 1 = + for all = 1, 2, ⋯ ,

if ±


if

+ 1,

where is a positive constant and one smaller than ).

ℎ ±

= =

= 0.

[10.24]

are positive or negative constants (but any

10.4. Experiment 10.4.1. General description The test columns are clamped into a solid holder, fixed to the concrete floor of the test dome at Harima Plant, Daicel Corporation (see Figure 10.3). The dome is licensed by the authorities for tests with rocket motors (pyrotechnic devices). The test columns are placed horizontally, and the vibrations are allowed only in the horizontal plane, such that gravity has no influence on the motions, as long as motions with small amplitude take place. The rocket motor is suspended from the ceiling by a long, thin wire. A strain gauge is mounted 70 mm from the clamped

214

Modern Trends in Structural and Solid Mechanics 1

end, on both sides of the column, to measure the axial strain, which is converted to an axial force. The lateral deflections are measured by an eddy current displacement sensor. For this purpose, a target disk is mounted 170 mm from the clamped end. Figure 10.6 shows a schematic of the test setup.

Figure 10.6. Schematic of the experimental setup (side view)

In order to protect the equipment from the, possibly very violent, fluttering motion of the test column, a couple of thin but strong and light-weighted wires of good length (harnesses for security) are loosely connected between the rocket motor and two holders on the floor (see Figure 10.3). 10.4.2. Rocket motor In the case of normal experiments with columns, for example under conservative loading, the test column is fixed, while the load is variable, increasing or decreasing. In the present rocket-applied experiment, a batch of identical rocket motors was specially prepared. Thus, the load is fixed, and the column design is variable. This means that the load, the rocket thrust in the present case, is specified first, and then the test columns are designed for test runs. The rocket motors used are of the solid-propellant type. They were specially designed and manufactured for the present experiment by Daicel Corporation (Harima Plant, Tatsuno, Japan). The casing has a total length (including mounting part) of 280 mm and a diameter of 103 mm, as shown in Figure 10.7. The mass of the empty motor casing is 3.4 kg. The rocket motor is filled with propellant with a mass of 1.1 kg. The total mass of a fresh motor is thus 4.5 kg, and it is assumed that

Shape-optimized Cantilevered Columns

215

the average mass of the motor is = 3.95 kg (= 3.4 + 1.1⁄2). The average mass moment of inertia around the center of gravity is = 3.01 x 10 kg m . The distance from the end of the column to the center of gravity of the motor is = 149.9 mm.

Figure 10.7. Sketch of the cross-section of the rocket motor:

① forward cap, ② casing, ③ propellant, ④ after cap, ⑤ nozzle

Figure 10.8. Thrust curve

216

Modern Trends in Structural and Solid Mechanics 1

The thrust curve is shown in Figure 10.8. The burnout time is approximately 4.5 s. During the first 0.13 s, the thrust increases from 0 to a peak value of 49.9 kgf (489 N). After the peak thrust point, the thrust decreases to a reasonably steady = 39.0 kgf (382 N). Based on experience from level, with the average value previous experiments (Sugiyama et al. 1995, 2000), the very short initial peak value is ignored, and the motor is assumed to produce a steady/nominal thrust of = 39.0 kgf (382 N). The tests described in this chapter were conducted on metric engineering units, where the unit of force is kilogram force (kgf). The unit of kgf can be converted into the SI force unit of Newton (N) by multiplication by the conversion factor ≈ 9.8 . 10.4.3. Columns The test pieces are beams cut out from an aluminum plate of thickness (“height” in the case of test beams) = 4 mm. The modulus of elasticity (Young’s modulus) is = 68.0 Gpa (obtained by the free vibration test in section 10.4.4). The density is = 2.71 x 10 kg m . The aim of the experiment is just to verify that a column designed by maximization of critical load under a constant volume is stable at a critical load higher than that for the initial column, and possibly stabilized by shape optimization with a targeted critical load less than that for the initial column, and with less volume. As only five rocket motors are presented, five tests can run. Under this restriction, the following missions are formulated: Mission I: using two motors for two test runs, the aim of the mission is to show that a column, designed by maximization of critical load under fixed volume (column no. 2), is stable while the initial column (column no. 1) is unstable. Mission II: using three motors for three test runs, the aim of the mission is to show that columns designed by minimization of volume under a fixed critical load (column nos. 4 and 5) may be stable, while the designed column (column no. 3) with less volume than that of the initial column is unstable. In these missions, five test runs are planned with the aim for each test run. And then, five test columns are designed to achieve the aims of the columns. It is noted that, as the applied load is constant and the length of the test columns is fixed, the variable in the experiment is the designed shape of the test columns, and so the design variable in the optimization is breadth . In the search for the designed shape, the finite element method is applied with 20 equal finite elements, and with 21

Shape-optimized Cantilevered Columns

217

design variables. Assuming that the nominal thrust is constant during loading and of magnitude = 39.0 kgf (382 N), five test columns are designed and machined for five test runs. Column no. 1: the aim of the first test run is to confirm that the initial column no. 1 becomes unstable by flutter. The critical flutter load for the initial column is designed as a load just less than the nominal thrust. Thus, the length of the initial column is determined as = 990 mm to have the critical load 38.2 kgf(374 N) < = 39.0 kgf (382 N). The length of the other four columns is fixed as = 990 mm. Column no. 1 is the uniform column having the length = 990 mm, the initial = 80 mm, and the height ℎ = 4 mm. The column gives the initial/ breadth reference column with the aspect ratio of the cross-section AR = b/h = 20.0. This gives a mass per unit length of = 0.866 kg m , and a radius of gyration of = ⁄ = 1.15 mm, where is the cross-sectional area and is the area moment of inertia. The slenderness ratio is given as = ⁄ = 861. The design for column nos. 2–5 is conducted under the constraint for the design variable with minimum allowable breadth, i.e. ≥ = ⁄2 = 40.0 mm. The minimum allowable breadth was determined from an engineering point of view. Vanishing minimum breadth seems obviously unrealistic. Most importantly, it is noted that the present optimum design is based on the assumption that the column is elastic during loading. If the cross-section vanishes, then the compressive stress on the cross-section at the vanishing breadth becomes infinite, going beyond the elastic limit of the column, and out of application of theory. Column no. 2: the aim of the second run is to confirm that the shape-optimized column, having the same volume as the initial column, is stable under the same thrust as that applied to column no. 1, validating the stabilizing effect of shape optimization by maximization of critical load. The design is obtained by application of [10.18]. The designed critical load is 43.9 kgf (430 N) > = 39.0 kgf (382 N), corresponding to an increase of 15 % over the initial design. The column is thus designed and expected to exhibit stable oscillations when being subjected to the average rocket thrust. The design of column nos. 3, 4 and 5 follows mission II. Thus, the shape optimization problem is defined as follows: mimimize volume (mass) subject to = fixed, by application of [10.19]. Column no. 3: the aim of the test is to show that the shape-optimized column under the constraint of fixed load with less volume is stable. The critical load is designed as 38.2 kgf (374 N) , exactly the same as for run no. 1, and the column is

218

Modern Trends in Structural and Solid Mechanics 1

thus predicted to be unstable. The volume is 87.4 % of the initial design’s volume, , aiming at a volume reduction of about 12 %. Column no. 4: the aim of the test is to validate the stabilizing effect of shape optimization under fixed load with less volume. The critical load was targeted to be 39.0 kgf (382 N) in the course of designing a test column with the load value coinciding with the flutter limit for test run no. 3. The resulting designed column has the critical load 39.1 kgf (383 N) and a volume of 89.35% of . The gain of volume reduction is about 10%. The stability of the vibration is expected to be critical under the applied load. Column no. 5: the aim of the test is to validate experimentally the stabilizing effect of shape optimization of the test column with higher critical load and less volume. The test column is targeted to have the critical load of 1.03 (40.0 kgf/ 392 N) and volume of 0.914 . The critical load is raised 2.6% and the volume is reduced by about 8.6%. The column is expected to remain stable under the applied load. Five test beams were prepared as shown in Figure 10.9. It is noted that the length of the naked beams, as shown in the figure, is 1138 mm. The length comprises of the clamp part of 100 mm, the column part of 990 mm, and the mounting part of 48 mm. It is interesting to see that there is a narrow and straight “corridor” between the clamped side and the forward portion in the cases of nos. 2–5, which is attributable = (1⁄2) = 40 mm. Geometrical data of to the minimum allowable breadth the four optimized columns, nos. 2–5, are listed in the appendix (section 10.8).

Figure 10.9. Test beams. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Shape-optimized Cantilevered Columns

a)

219

b)

Figure 10.10. Load–frequency curves: (a) for the initial column no. 1 and (b) for the optimized column no. 2 with the maximum critical load

Figure 10.10(a) shows the load–frequency curves for the initial column no. 1, while Figure 10.10(b) shows the curves for the optimized column no. 2. It can be observed that there are no significant changes in the development of these curves, contrary to cases with small or vanishing end mass (Langthjem and Sugiyama 1999b). The curves are similar for any of the five test columns. In addition, any of the optimized column nos. 2, 3, 4 and 5 satisfy the optimality criterion [10.24]. The designed columns are summarized in Table 10.1. Column no.

Volume

No. 1 No. 2

Designed critical thrust (kgf)

Predicted stability

38.2 (0.98 )

Flutter

43.9 (1.13 )

Stable

No. 3

0.874

37.2 (0. 95 )

Flutter

No. 4

0.894

39.1 (1. 00 )

Critical

No. 5

0.914

40.0 (1. 03 )

Stable

Table 10.1. Columns for five test runs

10.4.4. Free vibration test Before the planned flutter test, free vibration tests with the five test beams having a total length of 1038 mm (no rocket motor at tip end) were conducted. The test beams were fixed with the same clamp as was used in the rocket motor tests. The free vibrations, caused by a disturbance given at the free end for the first mode, were recorded to find the first natural eigenfrequency. The eigenfrequencies were measured three times and their average was taken to be the nominal values of the

220

Modern Trends in Structural and Solid Mechanics 1

test results. The test results are compared with the numerical predictions by the finite element method with 20 segments, as shown in Table 10.2. Column no.

No. 1

No. 2

No. 3

No. 4

No. 5

Numerical predictions

3.01

2.50

2.56

2.55

2.54

Test results

3.00

2.53

2.55

2.54

2.57

Table 10.2. First natural eigenfrequencies (Hz) of the five test beams

Young’s modulus (modulus of elasticity) was determined by the free vibration test of the reference beam. The first eigenfrequency of a uniform ⁄ , where is the length, is cantilevered beam is given by ≈ (0.560⁄ ) Young’s modulus, is the area moment of inertia, is the mass per unit length, and is the cross-sectional area. The eigenfrequency was measured by using the laser tracing system. Measurements were done on two different but identical uniform beams (for test beam no. 1), and on three different measure points on each beam. All measurements were repeated three times, giving 18 test cases in total. These tests gave = 3.23 (±0.006)Hz, and then = 68.0 Gpa was obtained. 10.5. Flutter test According to mission I, the first two test runs with column nos. 1 and 2 were conducted. Run no. 1: the vibrations of column no. 1 became unstable, as predicted, but in a much more drastic way than expected in the planned plot. In fact, column no. 1 was twisted up, breaking the harness wires (see Figures 10.3 and 10.11), in several rounds about the clamp, as shown in Figure 10.11 (a). Figure 10.12 (a) shows the dynamic response, i.e. the lateral deflection, observed in run no. 1. The response was measured 170 mm from the clamped end (see Figures 10.3 and 10.4). Run no. 2: the vibrations of column no. 2 became unstable, although they were predicted to remain stable in the design of the column. Figure 10.11 (b) shows the damaged column of test run no. 2. It is interesting to see that the narrow straight “corridor”, the weakest portion of the test column, remains straight, recalling that the weakest portion results in the heaviest bend in the case of conservative loading for buckling. Figure 10.12 (b) shows the response curve in run no. 2. The deflection turned on one side and then on the opposite side showing that the motion was not in divergence, but in flutter.

Shape-optimized Cantilevered Columns

a)

221

b)

Figure 10.11. Damaged test columns: (a) damaged column no. 1 and (b) damaged column no. 2. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

a)

b)

Figure 10.12. Dynamic responses in test runs nos. 1 and 2: (a) response in run no. 1 (the reading is only correct until the point “A”) and (b) response in run no. 2

From careful observation and comparison of the two dynamic responses (Figure 10.12), it may be seen, at least qualitatively, that the vibrations are stabilized by the optimal material distribution. In this sense, the experiment with the two test runs was partially successful in verifying the stabilizing effect of shape optimization; however, it must be said that the experiment was not completely successful with respect to a quantitative verification. The two test runs showed that the theoretical predictions of the critical length for the uniform column (and likewise for the optimally designed column as well) were not adequate, and that the actual critical length would be shorter than that predicted. Accordingly, planned test runs nos. 3, 4 and 5 were called off. Instead, it was decided to use the remaining three rocket motors to check the critical length of a uniform column, as used in run no. 1.

222

Modern Trends in Structural and Solid Mechanics 1

Another three extra test runs were conducted with uniform columns with decreasing length, shorter than the initial length of 990 mm. Run no. 6: the uniform column of length 900 mm was used to find that the vibrations still became unstable. The dynamic response is shown in Figure 10.13(a). It is observed by the curve in Figure 10.13(a) that the column loses stability by vibrations with increasing amplitude, i.e. by flutter. Run no. 7: the uniform column of length 850 mm was used to find the vibrations to remain stable. The dynamic response is shown in Figure 10.13(b). Run no. 8: the uniform column of length 875 mm, the mean value of the two lengths employed above, was used to find that the vibrations went unstable. The response is shown in Figure 10.13(c). The result of the extra runs is summarized graphically in Figure 10.14. The theoretical stability boundary is drawn with a full curve. The dots represent the test results. Stable and unstable vibrations are depicted by open and solid marks, respectively. Accordingly, the critical length of the uniform column exists in the < 875 mm. The mean value = 863 mm may be range of 850 mm ≤ employed as the critical length of the uniform column. If one takes the value of = 863 mm as the (estimated) critical length for the uniform column, the length is 13% shorter than the initial length of 990 mm, which was predicted for the first series of test runs.

a)

b)

c) Figure 10.13. Dynamic responses in the extra test runs with uniform columns: (a) response in run no. 6, (b) response in run no. 7 and (c) response in run no. 8

Shape-optimized Cantilevered Columns

223

Figure 10.14. Stability chart for the extra runs

10.6. Concluding remarks Optimum design of Beck’s column has been studied almost continuously during the last four decades. This chapter is one of the rare attempts to provide experimental validation of the stabilizing effect of shape optimization of slender structures subjected to follower loads (Borglund 1998, 1999). However, this chapter could not give a definite experimental verification of the stabilizing effect of shape optimization, as the theoretical prediction for the critical length failed to match the dynamic responses of the columns under a rocket-based follower force. Then, was the theory wrong? Or was the experiment done inappropriately? It is noted that the experiment has demonstrated some physical realities nonetheless, while the theory could not predict the dynamic response correctly. The stability analysis in this chapter has been developed based on the Euler–Bernoulli beam theory, assuming that the beam is sufficiently slender. However, the present test columns were prepared with plate-like (deep) beams, having a rather wide breadth of 80 mm and a height of 4.0 mm, i.e. the aspect ratio of the cross-section = ⁄ℎ = 20.0. It is recalled that the earlier three experiments (Sugiyama et al. 1995a, 1995b, 2000, 2019) were done successfully by using slender columns with aspect ratios of the cross-section in the range of 3.3 ≤ ≤ 5.0. This advised the authors to return to test columns with proper (as beams) sizes of the initial breadth of 30 mm and the height of 7 mm, with the aspect ratio = ⁄ℎ = 4.3 ≪ 20.0. The authors did so and succeeded in gaining good experimental verification of the stabilizing effect of shape optimization of a generalized version of Beck’s column (Sugiyama et al. 2012). The plate-like test columns were applied for the present experiment as a result of the authors’ intention to magnify the glamorous optimal shape design. The authors

224

Modern Trends in Structural and Solid Mechanics 1

proceeded from the successful free vibration tests, where the test results of the first eigenfrequency were well predicted by numerical analysis, to the somewhat unsuccessful flutter tests. However, it was revealed by the present experiment that the dynamic responses of the plate-like columns under a follower compressive loading were not predicted so as in the case of the free vibration test with deep beams, where the beams were free of an axial compressive loading. The present experiment suggests, in one more lesson, that there may be a gray zone between the beam theory and the plate theory, specifically in the case of deep beams under compressive loading. This chapter might eventually step into this zone. The gray zone remains to be investigated further in the future. For beams having plate-like form, Bolotin discussed in his book (Bolotin 1963, pp. 66–72, 104–119) the stability of the plane form of bending under the action of a transverse load in its plane of maximum stiffness. He suggested that the dynamics of the plane form can be described by two displacements, the lateral displacement and the angle of torsion (around the longitudinal axis ). The dynamic responses in the present experiment might be a kind of coupled bending–torsion flutter. This suggests to us that we might be better to study the dynamic stability of plate-like columns under a follower loading by considering both lateral and torsional motion. 10.7. Acknowledgments The authors would like to express their gratitude to Mr. M. Kobayashi (presently at Kawasaki Heavy Industries, Ltd) and Mr. T. Iwama (presently at Toyota Motor Corporation), former master students at the Sugiyama Laboratory, Department of Aerospace Engineering, Osaka Prefecture University; and Mr. H. Yutani at the Harima Plant, Daicel Corporation, who partly joined the present research project. 10.8. Appendix Designed breadth (mm) of the optimized column numbers 2, 3, 4 and 5 (height ℎ = 4.0 mm and length = 990.0 mm) Station (mm)

No. 2

No. 3

No. 4

No. 5

0.0

109.5

93.4

95.9

98.4

49.5

99.6

85.1

87.3

89.6

99.0

86.1

73.5

75.4

77.4

148.5

71.5

61.3

62.9

64.5

198.0

48.0

40.0

40.8

42.1

247.5

40.0

40.0

40.0

40.0

Shape-optimized Cantilevered Columns

Station (mm)

No. 2

No. 3

No. 4

No. 5

297.0

40.0

40.0

40.0

40.0

346.5

40.0

40.0

40.0

40.0

396.0

40.0

40.0

40.0

40.0

445.5

58.6

48.6

50.3

51.8

495.0

75.3

65.3

66.9

68.5

544.5

88.3

76.0

78.8

79.0

594.0

98.5

85.0

87.2

89.3

643.5

105.6

91.2

93.5

225

95.8

893.0

109.9

94.9

97.3

99.7

742.5

111.2

96.0

98.5

100.9

792.0

109.5

94.6

97.0

99.4

841.5

104.8

90.6

92.9

95.1

891.0

97.0

83.8

85.9

88.0

940.5

85.8

74.2

76.1

77.9

990.0

71.3

61.7

63.3

64.8

10.9. References Beck, M. (1952). Die Knicklast des einseitig eingespannten, tangential gedrückten Stabes. Z. Angew. Math. Phys., 3, 225–228. Bigoni, D. and Kirillov, O. (2019). Dynamic Stability and Bifurcation in Nonconservative Mechanics. Springer, Switzerland. Bolotin, V.V. (1963). Nonconservative Problems of the Theory of Elastic Stability. Pergamon Press, New York. Bolotin, V.V. and Zhinzher, N.I. (1969). Effects of damping on stability of elastic systems subjected to nonconservative force. Int. J. Solids Struct., 5, 965–989. Borglund, D. (1998). On the optimal design of pipes conveying fluid. J. Fluids Struct., 12, 353–365. Borglund, D. (1999). Active nozzle control and integrated design optimization of a beam subjected to fluid-dynamic forces. J. Fluids Struct., 12, 353–365. Claudon, J.-L. (1975). Characteristic curves and optimum design of two structures subjected to circulatory loads. J. Méca., 14, 531–543. Elishakoff, I. (2005). Controversy associated with the so-called follower forces. Appl. Mech. Rev., 58, 117–142.

226

Modern Trends in Structural and Solid Mechanics 1

Gutkowski, W., Mahrenholt, O., Pyrz, M. (1993). Minimum weight design of structures under nonconservative forces. In Optimization of Large Structural Systems, Rozvany, G.I.N. (ed.). Kluwer, Dordrecht. Hanaoka, H. and Washizu, K. (1980). Optimum design of Beck’s column. Comp. Struct., 11, 473–480. Herrmann, G. and Jong, I.-C. (1965). On the destabilizing effect of damping in nonconservative elastic systems. ASME J. Appl. Mech., 32, 592–597. Ishida, R. and Sugiyama, Y. (1995). Proposal of constructive algorithm and discrete shape design of the strongest column. AIAA J., 33(3), 401–406. Kirillov, O. (2013). Nonconservative Stability Problems of Modern Physics. De Gruyter, Berlin. Langthjem, M.A. and Sugiyama, Y. (1999a). Optimum design of Beck’s column with a constraint on the static buckling load. Struct. Optim., 18, 228–235. Langthjem, M.A. and Sugiyama, Y. (1999b). Optimum shape design against flutter of a cantilevered column with an end-mass of finite size subjected to a non-conservative load. J. Sound Vib., 226, 1–23. Langthjem, M.A. and Sugiyama, Y. (2000a). Dynamic stability of columns subjected to follower loads: A survey. J. Sound Vib., 238(5), 809–851. Langthjem, M.A. and Sugiyama, Y. (2000b). Optimum design of cantilevered columns under the combined action of conservative and nonconservative forces, Part I: The undamped case. Comp. Struct., 74, 385–398. Langthjem, M.A. and Sugiyama, Y. (2000c). Optimum design of cantilevered columns under the combined action of conservative and nonconservative forces, Part II: The damped case. Comp. Struct., 74, 399–408. Odeh, F. and Tadjbakhsh, I. (1975). The shape of the strongest column with a follower force. J. Optim. Theo. Applicat., 15, 103–108. Pedersen, P. and Seyranian, A.P. (1983). Sensitivity analysis for problems of dynamic stability. Int. J. Solids Struct., 19, 315–335. Plaut, R.H. (1972). Determining the nature of instability in nonconservative problems. AIAA J., 10, 967–968. Plaut, R.H. (1975). Optimal design for stability under dissipative, gyroscopic, or circulatory loads. In Optimization in Structural Design, Sawczuk, A. and Mroz, Z. (eds). SpringerVerlag, Berlin. Press, W.H., Teukolsky, S.A., Vetterling, V.T., Flannery, B.P. (1992). Numerical Recipes in Fortran, 2nd edition. Cambridge University Press, Cambridge. Ringertz, U.T. (1994). On the design of Beck’s column. Struct. Optim. (Struct. Multidisci. Optim.), 8, 120–124.

Shape-optimized Cantilevered Columns

227

Ryu, B.-J. and Sugiyama, Y. (2000). Dynamic stability of cantilevered Timoshenko columns subjected to a rocket thrust. Comp. Struct., 51, 331–335. Sugiyama, Y., Katayama, K., Kinoi, S. (1995a). Flutter of cantilevered column under rocket thrust. ASCE J. Aeros. Eng., 8, 9–15. Sugiyama, Y., Matsuike, J., Ryu, B.-J., Katayama, K., Kinoi, S., Enomoto, N. (1995b). Effect of concentrated mass on stability of cantilevers under rocket thrust. AIAA J., 33, 499–503. Sugiyama, Y., Katayama, K., Kiriyama, K., Ryu, B.-J. (2000). Experimental verification of dynamic stability of vertical cantilevered columns subjected to a sub-tangential force. J. Sound Vib., 236, 193–207. Sugiyama, Y., Langthjem, M.A., Iwama, T., Kobayashi, M., Katayama, K., Yutani, H. (2012). Shape optimization of cantilevered columns subjected to a rocket-based follower force and its experimental verification. Struct. Multidisc. Optim., 46, 826–838. Sugiyama, Y., Langthjem, M.A., Katayama, K. (2019). Dynamic Stability of Columns under Nonconservative Forces: Theory and Experiment. Springer, Switzerland. Tada, Y., Matsumoto, R., Oku, A. (1988). Shape determination of nonconservative structural systems (determination of optimum shape with stable critical load). In Computer Aided Optimum Design of Structures: Recent Advances, Hernández, S., Kassab, A.J., Brebbia, C.J. (eds). WIT Press, Southampton, UK. Tadjbakhsh, I. and Keller, J.B. (1962). Strongest columns and isoperimetric inequalities for eigenvalues. ASME J. Appl. Mech., 29, 159–164. Tezak, E.G., Nayfeh, A.H., Mook, D.T. (1982). Parametrically excited nonlinear multi-degree-of-freedom systems with repeated frequencies. J. Sound Vib., 85, 459–472. Venkateswara Rao, G. and Singh, G. (2001). Revisit to the stability of a uniform cantilever column subjected to Euler and Beck loads – Effect of realistic follower forces. Indian J. Eng. Mate. Sci., 8, 123–128. Ziegler, H. (1952). Die Stabilitätskrierien der Elastomechnik. Ing.-Archiv, 20, 45–56.

11 Hencky Bar-Chain Model for Buckling Analysis and Optimal Design of Trapezoidal Arches

This paper presents a Hencky bar-chain model (HBM) for buckling analysis and optimal design of funicular trapezoidal arches. The HBM comprises rigid beam segments connected by frictionless hinges and elastic rotational springs with stiffness C = EI/a, where EI is the member flexural rigidity and a is the segmental length. Based on the energy approach, canonical elastic and geometric stiffness matrices are derived for the HBM trapezoidal arch. The compatibility condition for trapezoidal arch shapes with general supports at different elevations is handled by using Lagrange multipliers. The eigenvalue equation for the buckling of trapezoidal arches is established by applying the principle of minimum potential energy. The lowest positive eigenvalue of the eigenvalue equation furnishes the buckling load. Two trapezoidal arch problems are solved to demonstrate the HBM for buckling analysis. Based on the developed HBM, we optimize the shape of symmetric trapezoidal arches of a prescribed volume of material for maximum buckling capacity. We consider trapezoidal arches under vertical point loads at different positions from the supports, as well as various base support conditions (such as hinged–hinged, fixed–hinged and fixed–fixed). The optimal arch height h and the cross-sectional area ratio AI/AH of the inclined members to the horizontal member are obtained, and a sensitivity analysis is performed to examine how the buckling load varies as the arch height changes, as well as how least-weight fully stressed arch solutions fare in comparison to the optimal arch designs against buckling.

Chapter written by Chien Ming WANG, Wen Hao PAN and Hanzhe ZHANG. Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

230

Modern Trends in Structural and Solid Mechanics 1

11.1. Introduction Funicular arches are elegant structures that transmit loads to the supports by purely compressive forces; there is no bending moment in the arch. In the design of funicular arches, it is important to check the arches against in-plane and out-of-plane buckling, which may cause premature failure. This paper focuses on funicular trapezoidal arches under two vertical point loads applied at the joints. In literature, some studies have been conducted on the stability of trapezoidal arches. Salem (1969) obtained the sway buckling load of a trapezoidal frame under axial compression. It was found that the elastic critical load increases as the inclination of columns or beams increases up to a maximum value, and then decreases again. The percentage increase in the critical load also depends on the bending stiffness of the frame members, i.e. the percentage increase is higher for hinged base frames than their fixed base counterparts. Wang (2007) investigated the in-plane buckling and nonlinear post-buckling of a trapezoidal frame with rigid members and flexible joints, where the horizontal member carries a rigid block. He also optimized the shape of the trapezoidal frame against buckling. Note that the deformation of the frame is only caused by the angle changes of four flexible connections whose stiffnesses are the same. Wang (2011) further studied the shape optimization problem of a similar trapezoidal frame against buckling by considering four flexible connections with different stiffnesses. Mohamadain et al. (2010) studied the stability behavior of trapezoidal frames exposed to elevated temperatures. They evaluated the effects of the column inclination, material degradation, support conditions and heating regimes of trapezoidal frames on their critical carrying capacities, and predicted the critical temperature of trapezoidal frames under design loads. Wang (2020) examined the stability and shape optimization problem of a trapezoidal vault with semi-rigid joints that has been subjected to hydrostatic pressure. It was found that the vault could have maximal volume and highest critical pressure when the top three rigid members were 39.64% of the base length forming a symmetrical trapezoid. The two upper joints should also be heavily strengthened. In this paper, we present the Hencky bar-chain model (HBM) for the in-plane elastic buckling analysis of funicular trapezoidal arches under two point loads. The three straight members of the trapezoidal arch may take on different sizes and the end supports may take on any elastically rotationally restrained condition. Based on the HBM, we investigate the optimal design of the funicular trapezoidal arch for maximum in-plane buckling capacity.

Hencky Bar-Chain Model for Buckling Analysis

231

11.2. Buckling analysis of trapezoidal arches based on the HBM Consider the buckling problem of a trapezoidal funicular arch under two vertical concentrated loads P and P at the member connections, as shown in Figure 11.1. The three inclined members have flexural rigidities EI1, EI2 and EI3, lengths L1, L2 and L3, and inclination angles 1, 2 and 3 with respect to the horizontal line. The supports at A and B are hinged supports with rotational spring stiffnesses KRA and KRB, respectively. When KRA = 0 and KRB = 0, we have hinged supports, and when KRA = ∞ and KRB = ∞, we have fixed supports. From a geometrical consideration, the inclination angle  of the sloping line passing through the two supporting points (see Figure 11.1) is given by

 L1 sin φ1 + L2 sin φ2 + L3 sin φ3   L  

γ = arctan 

[11.1]

where the horizontal span length L is given by

L = L1 cos φ1 + L2 cos φ2 + L3 cos φ3

[11.2]

φ3−3π/2 π/2+φ2

N2

αP

φ1-φ2

π/2-φ2

P

N1

EI2, L2

G2

φ2

π/2-φ1

N3

φ2-φ3+2π N2 EI3, L3

G1

φ3

EI1, L1 KRB B KRA A

φ1

γ L

Figure 11.1. Trapezoidal arch under two vertical concentrated loads. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

232

Modern Trends in Structural and Solid Mechanics 1

Based on the equilibrium of forces at the two joints G1 and G2, the axial forces in the three inclined members are given by

N1 =

P cos φ2 sin(φ1 − φ2 )

[11.3a]

N2 =

α P cos φ3 P cos φ1 , or N 2 = sin(φ1 − φ2 ) sin(φ2 − φ3 )

[11.3b]

N3 =

α P cos φ2 sin(φ2 − φ3 )

[11.3c]

Equation [11.3b] gives the following condition for the funicular shape of the trapezoidal arch:

α P cos φ3 sin(φ2 − φ3 ) cos φ1 P cos φ1 = α = sin(φ1 − φ2 ) sin(φ2 − φ3 ) sin(φ1 − φ2 ) cos φ3

[11.4]

Equation [11.4] implies that the load ratio α is fixed when the member inclination angles 1, 2 and 3 are prescribed. Alternatively, for a specified load ratio α and any of the two inclination angles ϕ, the remaining angle ϕ is fixed. 11.2.1. Description of the HBM

The Hencky bar-chain model (HBM) comprises a finite number of rigid segments connected by frictionless hinges and rotational springs. The HBM was proposed by Hencky (1920) based on an idea to replace deflection curves with piecewise linear curves. Since then, many researchers have been using this model for the analysis of beams and frames. These researchers have called the HBM by other names, such as the discrete element model (Sun et al. 1995), the segmented/articulated column (Wang 2001), the discrete link–spring model (Krishna and Ram 2007) and the microstructured beam model (Zhang et al. 2013). Silverman (1951) highlighted that the HBM is actually equivalent to the first-order central finite difference model (FDM), if the segmental length of the HBM is made equal to the nodal spacing of the FDM. Based on this equivalence, Wang et al. (2015) derived the boundary spring stiffnesses of the Hencky bar chain for elastic member-end restraints. Continuing along this line of investigation, Ruocco et al. (2016), Wang et al. (2016) and Zhang et al. (2016) developed the Hencky bar chain for beams with internal elastic restraints, with allowance for self-weight and varying cross-sections along the length. Zhang et al. (2017) and Pan et al. (2019a, 2019b)

Hencky Bar-Chain Model for Buckling Analysis

233

developed the Hencky bar chain for the buckling analysis of frames and circular arches, respectively. The HBM for buckling analysis of general funicular arches was further presented in the book by Wang et al. (2020b). The problem at hand is to determine the elastic buckling load P of the trapezoidal arch using the HBM. The HBM of the considered trapezoidal arch has each of its three members divided into n rigid segments of equal length, i.e. a1 = L1 n for the first member, a2 = L2 n for the second member and a3 = L3 n for the third member (see Figure 11.2). The rigid segments are connected by frictionless hinges with elastic rotational stiffnesses C1 = EI1 a1 , C2 = EI 2 a2 and C3 = EI 3 a3 for the first, second and third members, respectively.

αP ψ2n

P

ψ2n-1

ψ2n+1

CRG2

ψn

ψn+1

ψn-1

ψn+2

C2

CRG1

C1

C3

ψ3n-1

ψ2

ψ3n CRB

ψ1 CRA Figure 11.2. HBM for the trapezoidal arch. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

The HBM rotational spring stiffnesses at the supporting points, CRA and CRB, are given by (Wang et al. 2015) CRA =

1 1 1 + 2C1 K RA

CRB =

1 1 1 + 2C3 K RB

[11.5]

Note that for hinged supports, CRA = CRB = 0 since KRA = KRB = 0, whereas for fixed supports, CRA = 2C1 and CRB = 2C3 since KRA = KRB = ∞.

234

Modern Trends in Structural and Solid Mechanics 1

The HBM rotational spring stiffnesses at the joints are given by (Pan et al. 2019a; Wang et al. 2020b) CRG1 =

2 1 1 + C1 C2

CRG2 =

2 1 1 + C2 C3

[11.6]

11.2.2. HBM stiffness matrix formulation

For the trapezoidal arch problem, the 3n-order basis deformation vector of T segment rotations is selected as Ψ = (ψ 1 ,  , ψ n , ψ n +1 ,  , ψ 2 n , ψ 2 n +1 ,  , ψ 3 n ) . The elastic strain energy U of the HBM is given by U=

1 T Ψ [ KU Ψ ] Ψ 2

[11.7]

where [KUΨ] is the elastic stiffness matrix associated with the basis deformation vector, which is given by C1 [ KU 0 ]    K C K = [ UΨ ]  2[ U0]    C K [ ] 3 U0   CRA  [ 0n − 2 ]   CRG1 −CRG1  −CRG1 CRG1  + [ 0n − 2 ]  CRG 2   −CRG 2    

−CRG 2 CRG 2

[ 0n − 2 ]

             CRB 

[11.8]

where [KU0] is a canonical elastic stiffness matrix for the internal rotational springs in one member:

Hencky Bar-Chain Model for Buckling Analysis

 1 −1   −1 2 −1    [ KU 0 ] =  −1  −1    −1 2 −1   −1 1 

235

[11.9]

The potential energy V, due to the axial compressive load P in the HBM, can be expressed as V =

1 P ⋅ Ψ T [ KV Ψ ] Ψ 2

[11.10]

where the geometric stiffness matrix [KVΨ] is given by  cos φ2  a1 [ I n ]   sin( φ φ ) − 1 2     cos φ1 a2 [ I n ] [ KV Ψ ] = −   [11.11] sin(φ1 − φ2 )     α cos φ2  a3 [ I n ] sin(φ2 − φ3 )  

The total potential energy function is given by Π = U +V =

1 T Ψ ( [ KU Ψ ] + P ⋅ [ KV Ψ ] ) Ψ 2

[11.12]

11.2.3. Governing equation considering compatibility conditions

By considering the deformation at the buckled state and the fact that supports A and B do not move from their positions during buckling, the compatibility condition for the basis deformation vector Ψ can be established from geometrical considerations as

0   K μ  Ψ =   0 

[11.13]

236

Modern Trends in Structural and Solid Mechanics 1

where the (2×3n)-order matrix  K μ  for considering the compatibility conditions is given by  a1 sin φ1 ⋅ [1  1] a2 sin φ2 ⋅ [1  1] a3 sin φ3 ⋅ [1  1]   K μ  =  a cos φ ⋅ [1  1] a cos φ ⋅ [1  1] a cos φ ⋅ [1  1] 1 2 2 3 3  1 

[11.14]

In order to obtain the governing equation for buckling, we minimize the total potential energy subject to the satisfaction of the compatibility conditions (which contains the boundary conditions). By using Lagrange multipliers, we transform the constrained extremum problem to an unconstrained one. So the augmented total potential energy Γ (or Lagrangian) is given by Γ ( Ψ , μ1 , μ 2 ) = Π ( Ψ ) + {μ1

μ 2 } ⋅  K μ  Ψ

[11.15]

where 1 and 2 are the Lagrange multipliers. The stability criterion corresponds to the minimum value of the Lagrangian (equation [11.15]), i.e. ∂Γ ( Ψ , μ1 , μ 2 ) ∂ ( Ψ , μ1 , μ 2 )

[ K ] + P ⋅ [ K ]  K  T   Ψ  UΨ VΨ  μ   μ  = 0 = {}    1    0 K μ      μ2 

[11.16]

In order to determine the buckling load of the arch, equation [11.16] has to take on a non-trivial solution of the vector (Ψ, 1, 2), which is achieved by setting the determinant of the governing stiffness matrix to zero, i.e. [ K ] + P ⋅ [ K ]  K  T  UΨ VΨ  μ =0 det    K μ  0  

[11.17]

The buckling load Pcr is obtained by solving this characteristic equation [11.17] for the lowest positive root. The eigenvector (Ψ, 1, 2) in equation [11.16], associated with Pcr, furnishes the buckling mode. 11.2.4. Verification of the HBM

Consider funicular trapezoidal arches under two loads P and P, as shown in Figure 11.1. Table 11.1 presents the prescribed parameters (1, 2, 3, L1, L2, L3, EI1,

Hencky Bar-Chain Model for Buckling Analysis

237

EI2, EI3) for the two trapezoidal arches considered. The span length Lh and the load factor  are calculated from equations [11.2] and [11.4], respectively. Table 11.2 presents the buckling load parameters PL2/(EI) for n-segment HBM trapezoidal arches for the different shapes. It can be seen that the buckling loads converge monotonically to their corresponding continuum solutions (n = ∞) from below with increasing segment numbers n. The buckling modes of the trapezoidal arches 1 and 2, modeled by the HBM with n = 8, are shown in Figures 11.3 and 11.4, respectively.

Arch shape

Span length L

Load factor 

Symmetric trapezoidal arch 1 with

Non-symmetric trapezoidal arch 2 with

1 = /3, 2 = 0, 3 = 5/3

1 = /4, 2 = 11/6, 3 = 5/3

EI1 = EI2 = EI3 = EI

EI1= EI3=2EI, EI2=EI

L1 = L2 = L3= L0

L1=2L0, L2=1.5L0, L3=L0

 π   5π cos  3  + cos ( 0 ) + cos  3     = 2 L0

   L0 

π  5π  sin  −  cos 3  3  =1 α= π 5π sin cos 3 3

π   11π   5π  L1 cos   + L2 cos   +L3 cos   4 6      3  = 3.213L0

π  11π 5π  − sin   cos 3  4  6 = 0.732 5π  π 11π  sin  − cos  6  3 4

Table 11.1. Parameters for non-symmetrical trapezoidal arches 1 and 2

n

Trapezoidal arch 1

Trapezoidal arch 2

4

38.784

68.647

8

40.314

71.352

16

40.705

72.032

32

40.804

72.202

∞*

40.828

72.260

*The continuum solution is obtained using the matrix stiffness method with stability functions (Yuan 2008).

Table 11.2. Buckling load parameters PL2/(EI) of trapezoidal arches – n-segment HBM solutions and continuum solutions

238

Modern Trends in Structural and Solid Mechanics 1

P P

Figure 11.3. Buckling mode of trapezoidal arch 1 modeled by the HBM

P

0.732P

Figure 11.4. Buckling mode of trapezoidal arch 2 modeled by the HBM

It can be seen that the HBM furnishes accurate buckling loads for the trapezoidal arches by taking a sufficient number of segments, say 20 segments per member. As the HBM is a discrete structural model, any local changes to the cross-section of the arch due to damage or stiffening can be readily accommodated by changing the rotational spring stiffnesses appropriately at that location. Also, semi-rigid joints and support conditions may be easily taken care of by changing the rotational spring stiffnesses at the joints and supporting points.

Hencky Bar-Chain Model for Buckling Analysis

239

11.3. Optimal design of symmetric trapezoidal arches 11.3.1. Problem definition

Consider the elastic in-plane buckling of a symmetric trapezoidal arch, carrying two equal vertical point loads P at the joints G1 and G2, as shown in Figure 11.5. The span length of the arch is L. The horizontal distances of the total span length L divided by P are b, bH and b. The height of the arch is h. P

P

LH G1

EIH, AH

G2

θ LI

EII AI

EII AI

h

LI

fixed or pinned A b

bH

b

B

L Figure 11.5. Trapezoidal arch

The inclined members have a flexural rigidity EII, cross-sectional area AI and length LI, while the horizontal member has a flexural rigidity EIH, cross-sectional area AH and length LH. Considering the members of a geometrically similar cross-section, the second moment of area I is related to the cross-sectional area A by I = A2

[11.18]

240

Modern Trends in Structural and Solid Mechanics 1

where the parameter  depends on the cross-sectional shape. For example,  = 1/(4) for a circular solid cross-section and  = 1/12 for a square solid section. The arch supporting points A and B are either hinged–hinged, fixed–fixed or fixed–hinged. The optimization problem at hand may be stated as follows: Maximize buckling load Pcr subjected to a prescribed arch volume V constraint, i.e. 2AILI + AHLH = V

[11.19]

By using this constraint, the decision variables involved in this optimization problem are the arch height h and the cross-sectional area ratio AI/AH of the two members. Both variables are free to take any positive value in the optimization procedure. 11.3.2. Optimization procedure

The buckling load (calculated from the HBM with 40 rigid segments per member) has to be maximized by selecting the appropriate arch height h and the cross-sectional area ratio AI/AH of the two members. The final unconstrained optimization problem, involving two decision variables (h and AI/AH), may be solved by using the Nelder and Mead simplex method (Nelder and Mead 1965), which does not require the evaluations of any gradients. 11.3.3. Optimal solutions

Figure 11.6 shows the variations of the optimal height hopt/L of the trapezoidal arch with respect to the horizontal distance b/L of the point load from the support, for various combinations of support conditions. It can be seen that the optimal height hopt/L increases from approximately 0.15 to 0.25, as the horizontal distance b/L varies from 0.15 to 0.4. The optimal arch solution with h/L = 0.15–0.25 is considered to be a high arch (Bažant and Cedolin 2010; Wang et al. 2020a) because there is no significant increase in axial forces in members during the deformation of the arch. Therefore, the bifurcation buckling due to the compressive axial forces in the members is the dominant behavior for the considered trapezoidal arch structure.

Hencky Bar-Chain Model for Buckling Analysis

241

Figure 11.6. Variations of optimal height of trapezoidal arches for different support conditions with respect to b/L. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Figure 11.7 shows the variations of the optimal cross-sectional area ratio AI/AH, with respect to the horizontal distance b/L of the point load from the support for different base support conditions. It can be seen that the optimal cross-sectional area ratio AI/AH varies from approximately 0.8 to 1.6, as the horizontal distance b/L varies from 0.15 to 0.4. Figure 11.8 shows the variations of the non-dimensional maximum buckling load Pmax L4 / α EV 2 , with respect to the horizontal position b/L of the load from the

(

)

support for the various base support conditions. For the fixed–hinged arch, the maximum buckling load lies within the buckling loads of the fixed–fixed and hinged–hinged arches.

242

Modern Trends in Structural and Solid Mechanics 1

Figure 11.7. Variations of optimal cross-sectional area ratio of trapezoidal arches for different base support conditions with respect to b/L. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Figure 11.8. Variations of the maximum buckling load of trapezoidal arches for different base support conditions with respect to b/L. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Hencky Bar-Chain Model for Buckling Analysis

243

11.3.4. Sensitivity analysis of optimal solutions

Consider a trapezoidal arch of a given volume of material and subject to two vertical point loads, whose horizontal position b/L from the support is prescribed. Figure 11.9(a)–(c) shows the sensitivity of the maximum buckling load for this arch with respect to the vertical height h/L. It can also be seen from Figure 11.9 that the optimal height for the maximum buckling load h/L varies from approximately 0.15 to 0.25. Based on this sensitivity analysis, it can be noted that the trapezoidal arch with a vertical height h/L that is non-optimal may have significantly smaller buckling loads than the corresponding optimal arch designed against buckling. Take fixed–fixed trapezoidal arches as an example. When b/L = 0.2, the trapezoidal arch with a height h/L = 0.4 has a buckling load that is only 47.3% of that associated with the optimal arch design against buckling (where h/L = 0.1678).

(a) Fixed–fixed arches

244

Modern Trends in Structural and Solid Mechanics 1

(b) Hinged–hinged arches

(c) Fixed–hinged arches Figure 11.9. Sensitivity of the maximum buckling load of trapezoidal arches with respect to h/L, for three base support conditions. For a color version of this figure, see www.iste.co.uk/challamel/mechanics1.zip

Hencky Bar-Chain Model for Buckling Analysis

245

11.3.5. Comparison with the buckling load of optimal fully stressed trapezoidal arches

The optimal design of fully stressed trapezoidal arches for least weight was investigated by Rozvany and Wang (1983). The shape of the optimal arch with base supports at the same level has to satisfy the unit mean square slope optimality condition, i.e. L

2

 dy  0  dx  dx = 1.0 L

[11.20]

where y is the elevation of the member above the horizontal supporting line. For a symmetrical trapezoidal arch under two equal vertical point loads, the above optimality condition yields

hopt =

bL 2

AI 1 + 2b L = 2b L AH

[11.21]

[11.22]

Such optimal fully stressed trapezoidal arches, whose shapes are described by equation [11.21], may have significantly smaller buckling loads than the corresponding optimal arch designed against buckling. For example, consider a hinged–hinged trapezoidal arch with loading at b/L = 0.3. According to equation [11.21], the optimal fully stressed arch has an optimal height h/L ≈ 0.39, and its buckling load is only 65% of that associated with the optimal arch that is designed against buckling (where h/L = 0.201). 11.4. Concluding remarks

In this chapter, we have presented the Hencky bar-chain model (HBM) for the elastic in-plane buckling analysis of a funicular trapezoidal frame, under two point loads at the joints between members. The supports may be hinged (defined by an HBM rotational spring stiffness of zero) or fixed (defined by an HBM rotational spring stiffness that is twice the HBM internal rotational spring stiffness of the member). The HBM furnishes accurate buckling solutions when a sufficient number of bar segments is adopted (say 20 segments) for each member.

246

Modern Trends in Structural and Solid Mechanics 1

Based on the developed HBM, optimal designs of trapezoidal arches against buckling were obtained for various load positions and base support conditions. As the horizontal distance b/L varies from 0.15 to 0.4, the optimal height hopt/L increases from approximately 0.15 to 0.25, and the optimal cross-sectional area ratio AI/AH varies from approximately 0.8 to 1.6. A sensitivity analysis of the maximum buckling load with respect to h/L indicates that the trapezoidal arch with a vertical height h/L that is non-optimal may have significantly smaller buckling loads than the corresponding optimal arch designed against buckling. Therefore, for an economical design of trapezoidal arches considering both strength and stability, the optimal arch solution against buckling should also be used in addition to the condition for the optimal fully stressed trapezoidal arches. 11.5. References Hencky, H. (1920). Über die angenäherte Lösung von Stabilitätsproblemen im Raum mittels der elastischen Gelenkkette. Der Eisenbau, 11, 437–452. Krishna, S.G. and Ram, Y.M. (2007). Discrete model analysis of optimal columns. International Journal of Solids and Structures, 44(22–23), 7307–7322. Mohamadain, M.A., El-Arabi, I.A., Nassef, W.M. (2010). Inelastic stability of steel trapezoidal frames at elevated temperatures. Ain Shams Engineering Journal, 1(1), 21–30. Nelder, J.A. and Mead, R. (1965). A simplex method for function minimization. The Computer Journal, 7(4), 308–313. Pan, W.H., Wang, C.M., Zhang, H. (2019a). Hencky bar-chain model for buckling analysis of non-symmetric portal frames. Engineering Structures, 182, 391–402. Pan, W.H., Wang, C.M., Zhang, H. (2019b). Matrix method for buckling analysis of frames based on Hencky bar-chain model. International Journal of Structural Stability and Dynamics, 19(8), 1950093. Rozvany, G.I.N. and Wang, C.M. (1983). On plane Prager-structures I. International Journal of Mechanical Sciences, 25(7), 519–527. Ruocco, E., Zhang, H., Wang, C.M. (2016). Hencky bar-chain model for buckling analysis of non-uniform columns. Structures, 6, 73–84. Salem, A.H. (1969). Buckling of trapezoidal frames permitted to sway. Journal of the Structural Division, 95(12), 2621–2640. Silverman, I.K. (1951). Discussion on the paper of “Salvadori M.G., Numerical computation of buckling loads by finite differences, Transactions of the ASCE, 116, 590–636, 1951”. Transactions of the ASCE, 116(1), 625–626. Sun, C., Shaw, W.J.D., Vinogradov, A.M. (1995). Discrete-element model for buckling analysis of thin ring confined within rigid boundary. Journal of Engineering Mechanics, 121(1), 71–79.

Hencky Bar-Chain Model for Buckling Analysis

247

Wang, C.Y. (2001). Stability of a heavy segmented column. Mechanics Research Communications, 28(5), 493–497. Wang, C.Y. (2007). Optimum trapezoidal frame with rigid members and flexible joints. International Journal of Non-linear Mechanics, 42(5), 760–764. Wang, C.Y. (2011). Stability and optimization of the portal frame with rigid members and flexible joints. Journal of Engineering Mechanics, 137(5), 366–370. Wang, C.Y. (2020). Stability and optimization of a pressure-loaded trapezoidal vault with semi-rigid joints. International Journal of Structural Stability and Dynamics, 20(1), 2071001. Wang, C.M., Zhang, H., Gao, R.P., Duan, W.H., Challamel, N. (2015). Hencky bar-chain model for buckling and vibration of beams with elastic end restraints. International Journal of Structural Stability and Dynamics, 15(7), 1540007. Wang, C.M., Zhang, H., Challamel, N., Xiang, Y. (2016). Buckling of nonlocal columns with allowance for selfweight. Journal of Engineering Mechanics, 142(7), 04016037. Wang, C.M., Pan, W.H., Zhang, J.Q. (2020a). Optimal design of triangular arches against buckling. Journal of Engineering Mechanics, 146(7), 04020059. Wang, C.M., Zhang, H., Challamel, N., Pan, W.H. (2020b). Hencky Bar-Chain/Net for Structural Analysis. World Scientific, Singapore. Yuan, S. (2008). Programming Structural Mechanics, 2nd edition. Higher Education Press, Beijing. Zhang, Z., Challamel, N., Wang, C.M. (2013). Eringen’s small length scale coefficient for buckling of nonlocal Timoshenko beam based on microstructured beam model. Journal of Applied Physics, 114(11), 114902. Zhang, H., Wang, C.M., Challamel, N. (2016). Buckling and vibration of Hencky bar-chain with internal elastic springs. International Journal of Mechanical Sciences, 119, 383–395. Zhang, H., Wang, C.M., Challamel, N. (2017). Small length scale coefficient for Eringen’s and lattice-based continualized nonlocal circular arches in buckling and vibration. Composite Structures, 165, 148–159.

List of Authors

Jacob ABOUDI Tel Aviv University Israel

Srinandan DASMAHAPATRA University of Southampton UK

Mariano P. AMEIJEIRAS CEFyN-UNC Argentina

Aharon DEUTSCH Technion – Israel Institute of Technology Israel

Romesh C. BATRA Virginia Polytechnic Institute and State University USA Atul BHASKAR University of Southampton UK Emanuela BOLOGNA Univeristy of Palermo Italy

Mario DI PAOLA University of Palermo Italy Moshe EISENBERGER Technion – Israel Institute of Technology Israel Uğurcan EROĞLU Izmir University of Economics Turkey

Devin BURNS Virginia Polytechnic Institute and State University USA

Rivka GILAT Ariel University Israel

Noël CHALLAMEL University of Southern Brittany France

Luis A. GODOY CEFyN-UNC Argentina

Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

250

Modern Trends in Structural and Solid Mechanics 1

Valery I. GULYAYEV National Transport University Kyiv Ukraine

Natalya V. SHLYUN National Transport University Kyiv Ukraine

Andrii IAKOVLIEV University of Southampton UK

Yoshihiko SUGIYAMA Osaka Prefecture University Japan

Julius KAPLUNOV Keele University UK

Izuru TAKEWAKI Kyoto University Japan

Kazuo KATAYAMA Daicel Corporation Japan

Joseph TENENBAUM Technion – Israel Institute of Technology Isreal

Mikael A. LANGTHJEM Aarhus University Denmark Wen Hao PAN Zhejiang University China Achille PAOLONE Sapienza University of Rome Italy Vincent PICANDET University of Southern Brittany France Giuseppe RUTA Sapienza University of Rome and National Group for Mathematical Physics Italy

Ekrem TÜFEKCI Istanbul Technical University Turkey Chien Ming WANG Queensland University Australia Hanzhe ZHANG Beijing Institute of Technology China Massimiliano ZINGALES University of Palermo Italy

Index

A, B, C

D, E, F

asymptotic approach, 67, 70, 86 Beck’s column, 43, 44, 54, 57, 60, 65, 201–205, 223 bending, 159, 161, 162, 174, 177, 178, 181–183, 188, 195, 203, 206, 224, 230 bifurcation, 122, 126–128, 132, 136–139, 159, 161, 164, 165, 168, 173, 178, 182, 196, 197, 240 boundary layer, 67, 69, 70, 72, 74, 81–86, 181 buckling, 17–20, 22, 23, 25–40, 91–94, 97–99, 105–108, 121, 122, 129, 136–139, 159–161, 173–175, 177–179, 182, 184, 185, 188, 190–193, 195–198, 205, 220, 229–231, 233, 235–246 inelastic, 204, 205 post-, 121, 122, 139, 174, 230 wavelength, 146, 148, 153, 154, 156 continuum solutions, 237 compliance, 133, 134, 138 crack, 121, 122, 133–137, 139 critical axial force, 195 load, 166, 173, 178, 182, 201–204, 209, 211, 216–219, 230 curvature, 187, 188, 192, 195, 203

damage, 121, 122, 136–139, 238 deep curvilinear borehole channels, 179 difference equation, 67, 69–72, 74–78 direct and indirect interactions, 67–71, 76, 77, 80–82, 84–86 discrete model, 232, 238 disordered chains, 31, 39, 40 drill string buckling, 177, 184, 195, 197, 198 dynamic instability, 159–161, 168–170, 174, 208 elastic, 1–6, 9, 15, 67–70, 72, 73, 77, 86, 161, 162, 175, 178, 179, 182, 183, 185, 187, 188, 192, 194, 201, 202, 216, 217, 220, 229, 230, 232–234, 239, 245 chains, 17 lattice, 68, 72, 73, 77 springs, 67, 139 elastically coupled rigid rods, 23, 31, 39, 40 equation, 1–7, 9, 10, 14, 15, 67–79, 83–85, 159, 164, 168–170, 179–185, 187–195, 197, 206–208, 229, 232, 235–237, 245 exact lattice solution, 84 solution, 8, 15, 69, 70, 91–93, 98

Modern Trends in Structural and Solid Mechanics 1: Statics and Stability, First Edition. Edited by Noël Challamel, Julius Kaplunov and Izuru Takewaki. © ISTE Ltd 2021. Published by ISTE Ltd and John Wiley & Sons, Inc.

252

Modern Trends in Structural and Solid Mechanics 1

experimental device, 201, 205, 206, 214, 218, 223 fiber-reinforced composite laminates, 1 finite difference model, 232 element method, 14, 91, 175, 208, 216, 220 flutter, 201–204, 206, 208, 209, 212–214, 217–220, 222, 224 follower load, 202–205, 223, 224 fractional -order viscoelastic foundation, 51 differential equation, 53, 55, 57, 59 Zener model, 44, 51, 53 G, H, I, J geometric imperfection, 163 geometrically nonlinear, 159, 164, 165, 174 gravity, 179, 184, 185, 187, 191, 193, 207, 213, 215 Hencky bar-chain model, 229–238, 240, 245, 246 hyperbolic functions, 95, 97, 116, 117 imperfection sensitive system, 161, 165, 174 jump conditions, 135 L, M, N, O Lagrange polynomials, 6 lattice, 67–73, 77, 79–82, 84–86 layered structures, 18, 19, 23, 39 localization, 18, 19, 22, 23, 31, 33–40, 67, 68, 70, 76, 81, 82, 86, 178, 180, 182, 194, 196–198 long-range interaction, 67–72 material hereditariness, 44, 45 non-conservative systems, 56 numerical integration, 7, 159, 164, 170, 183, 190, 192, 194, 197 optimization, 201–205, 209, 211, 216–219, 221–224, 229, 230, 240

P, R parabolic arches, 121, 122, 130, 132, 135, 137–139 periodic systems, 18 perturbation approach, 126, 139 plasticity, 146, 151, 153–156, 175 plates, 2–4, 6, 8–15, 91–94, 96–109, 111, 112, 173 rocket motor, 201, 203, 204, 206, 207, 213–216, 219, 221 rotational stiffness, 71, 229, 231, 233, 234, 238, 245 Routh–Hurwitz theorem, 43, 44, 60–62, 65 S, T sequential linear optimization, 211 shear buckling mode, 145 singularly perturbed problems, 177, 181, 183, 197 stability/instability, 17–20, 22, 23, 26, 28, 34–36, 38–40, 68, 91, 159–161, 164–170, 172–175, 177, 178, 184, 188, 192, 195, 197, 198, 201, 202, 204–206, 208, 216–224, 230, 236, 237, 246 static behavior, 67, 69, 70, 86, 159, 161, 164 statics, 69, 122 stress intensity factors, 133 three-dimensional exact solution, 8 linear elasticity equations, 2, 15 transfer matrix method, 132 transverse buckling mode, 145 strain, 5 stess, 1, 2, 5, 15 trapezoidal arches, 229–234, 236–246 triangular impulse, 164 two degree-of-freedom system, 159, 161, 162, 164–168, 174

Index

U, V, W uniform load, 71, 77, 79–82, 85 vibration test, 216, 219, 220, 224 volume fraction, 2, 145, 151–156 minimization under constant critical load, 202, 203, 209 wave propagation, 68, 146, 148, 152, 153, 156

253

Summary of Volume 2

Preface Noël CHALLAMEL, Julius KAPLUNOV and Izuru TAKEWAKI Chapter 1. Bolotin’s Dynamic Edge Effect Method Revisited (Review) Igor V. ANDRIANOV and Lelya A. KHAJIYEVA 1.1. Introduction 1.2. Toy problem: natural beam oscillations 1.3. Linear problems solved 1.4. Generalization for the nonlinear case 1.5. DEEM and variational approaches 1.6. Quasi-separation of variables and normal modes of nonlinear oscillations of continuous systems 1.7. Short-wave (high-frequency) asymptotics. Possible generalizations of DEEM 1.8. Conclusion: DEEM, highly recommended 1.9. Acknowledgments 1.10. Appendix 1.11. References Chapter 2. On the Principles to Derive Plate Theories Marcus AßMUS and Holm ALTENBACH 2.1. Introduction 2.2. Some historical remarks 2.3. Possibilities to formulate plate theories 2.3.1. Theories based on hypotheses 2.3.2. Reduction of the governing equations by mathematical techniques 2.3.3. Direct approach 2.3.4. Consistent approach

Modern Trends in Structural and Solid Mechanics

2.4. Shear correction 2.5. Conclusion 2.6. References Chapter 3. A Softening–Hardening Nanomechanics Theory for the Static and Dynamic Analyses of Nanorods and Nanobeams: Doublet Mechanics

Ufuk GUL and Metin AYDOGDU 3.1. Introduction 3.2. Doublet mechanics formulation 3.3. Governing equations 3.3.1. Static equilibrium equations of a nanorod with periodic micro- and nanostructures 3.3.2. Equations of motion of a nanorod with periodic micro- and nanostructures 3.3.3. Static equilibrium equations of a nanobeam with periodic micro- and nanostructures 3.3.4. Equations of motion of a nanobeam with periodic micro- and nanostructures 3.4. Analytical solutions 3.4.1. Axial deformation of nanorods with periodic nanostructures 3.4.2. Vibration analysis of nanorods with periodic nanostructures 3.4.3. Axial wave propagation in nanorods with periodic nanostructures 3.4.4. Flexural deformation of nanobeams with periodic nanostructures 3.4.5. Buckling analysis of nanobeams with periodic nanostructures 3.4.6. Vibration analysis of nanobeams with periodic nanostructures 3.4.7. Flexural wave propagation in nanobeams with periodic nanostructures 3.5. Numerical results 3.6. Conclusion 3.7. References Chapter 4. Free Vibration of Micro-Beams and Frameworks Using the Dynamic Stiffness Method and Modified Couple Stress Theory J.R. BANERJEE 4.1. Introduction 4.2. Formulation of the potential and kinetic energies 4.3. Derivation of the governing differential equations 4.4. Development of the dynamic stiffness matrix 4.4.1. Axial stiffnesses 4.4.2. Bending stiffnesses 4.4.3. Combination of axial and bending stiffnesses 4.4.4. Transformation matrix

Summary of Volume 2

4.5. Application of the Wittrick–Williams algorithm 4.6. Numerical results and discussion 4.7. Conclusion 4.8. Acknowledgments 4.9. References Chapter 5. On the Geometric Nonlinearities in the Dynamics of a Planar Timoshenko Beam Stefano LENCI and Giuseppe REGA 5.1. Introduction 5.2. The geometrically exact planar Timoshenko beam 5.3. The asymptotic solution 5.4. The importance of nonlinear terms 5.4.1. An initial case 5.4.2. The effect of the slenderness 5.4.3. The effect of the end spring 5.4.4. The effect of the resonance order 5.5. Simplified models 5.5.1. Neglecting axial inertia 5.5.2. One-field equation 5.5.3. The Euler–Bernoulli nonlinear beam 5.6. Conclusion 5.7. References Chapter 6. Statics, Dynamics, Buckling and Aeroelastic Stability of Planar Cellular Beams Angelo LUONGO 6.1. Introduction 6.2. Continuous models of planar cellular structures 6.2.1. Timoshenko beam 6.2.2. Shear beam 6.2.3. Elastic constant identification 6.3. The grid beam 6.3.1. Rigid transverse model 6.3.2. Flexible transverse model 6.3.3. Comparison among models 6.4. Buckling 6.4.1. Formulation 6.4.2. Critical loads 6.5. Dynamics 6.5.1. Timoshenko beam 6.5.2. Shear beam and discrete spring–mass model

Modern Trends in Structural and Solid Mechanics

6.6. Aeroelastic stability 6.6.1. Modeling a base-isolated tower 6.6.2. Critical wind velocity 6.7. References Chapter 7. Collapse Limit of Structures under Impulsive Loading via Double Impulse Input Transformation

Izuru TAKEWAKI, Kotaro KOJIMA and Sae HOMMA 7.1. Introduction 7.2. Collapse limit corresponding to the critical timing of second impulse 7.3. Classification of collapse patterns in non-critical case 7.4. Analysis of collapse limit using energy balance law 7.4.1. Collapse Pattern 1’ 7.4.2. Collapse Pattern 2’ 7.4.3. Collapse Pattern 3’ 7.4.4. Collapse Pattern 4’ 7.5. Verification of proposed collapse limit via time-history response analysis 7.6. Conclusion 7.7. References Chapter 8. Nonlinear Dynamics and Phenomena in Oscillators with Hysteresis Fabrizio VESTRONI and Paolo CASINI 8.1. Introduction 8.2. Hysteresis model and SDOF response to harmonic excitation 8.3. 2DOF hysteretic systems 8.3.1. Equations of motion 8.3.2. Modal characteristics 8.4. Nonlinear modal interactions in 2DOF hysteretic systems 8.4.1. Top-hysteresis configuration (TC) 8.4.2. Base-hysteresis configuration (BC) 8.5. Conclusion 8.6. Acknowledgments 8.7. Appendix: Mechanical characteristics of SDOF and 2DOF systems 8.8. References Chapter 9. Bridging Waves on a Membrane: An Approach to Preserving Wave Patterns Peter WOOTTON and Julius KAPLUNOV 9.1. Introduction 9.2. Problem statement

Summary of Volume 2

9.3. Homogenized bridge 9.4. Internal reflections 9.5. Discrete bridge 9.6. Net bridge 9.7. Concluding remarks 9.8. Acknowledgments 9.9. References Chapter 10. Dynamic Soil Stiffness of Foundations Supported by Layered Half-Space Yang Z HOU and Wei-Chau X IE 10.1. Introduction 10.2. Generation of dynamic soil stiffness 10.2.1. Dynamic stiffness matrix under point loads 10.2.2. Formulation of the flexibility function 10.2.3. Formulation of Green’s influence function 10.2.4. Total dynamic soil stiffness by the boundary element method 10.3. Numerical examples of the generation of dynamic soil stiffness 10.3.1. A rigid square foundation supported by a layer on half-space 10.3.2. A rigid circular foundation supported by a layer on half-space 10.3.3. A rigid circular foundation supported by half-space and a layer on half-space 10.4. Numerical examples of the generation of FRS 10.5. Conclusion 10.6. References

Summary of Volume 3

Preface Noël CHALLAMEL, Julius KAPLUNOV and Izuru TAKEWAKI Chapter 1. Optimization in Mitochondrial Energetic Pathways Haym BENAROYA 1.1. Optimization in neural and cell biology 1.2. Mitochondria 1.3. General morphology; fission and fusion 1.4. Mechanical aspects 1.5. Mitochondrial motility 1.6. Cristae, ultrastructure and supercomplexes 1.7. Mitochondrial diseases and neurodegenerative disorders 1.8. Modeling 1.9. Concluding summary 1.10. Acknowledgments 1.11. Appendix 1.12. References Chapter 2. The Concept of Local and Non-Local Randomness for Some Mechanical Problems Giovanni FALSONE and Rossella LAUDANI 2.1. Introduction 2.2. Preliminary concepts 2.2.1. Statically determinate stochastic beams 2.2.2. Statically indeterminate stochastic beams

Modern Trends in Structural and Solid Mechanics

2.3. Local and non-local randomness 2.3.1. Statically determinate stochastic beams 2.3.2. Statically indeterminate stochastic beams 2.3.3. Comments on the results 2.4. Conclusion 2.5. References Chapter 3. On the Applicability of First-Order Approximations for Design Optimization under Uncertainty Benedikt KRIEGESMANN 3.1. Introduction 3.2. Summary of first- and second-order Taylor series approximations for uncertainty quantification 3.2.1. Approximations of stochastic moments 3.2.2. Probabilistic lower bound approximation 3.2.3. Convex anti-optimization 3.2.4. Correlation of probabilistic approaches and convex anti-optimization 3.3. Design optimization under uncertainty 3.3.1. Robust design optimization 3.3.2. Reliability-based design optimization 3.3.3. Optimization with convex anti-optimization 3.4. Numerical examples 3.4.1. Imperfect von Mises truss analysis 3.4.2. Three-bar truss optimization 3.4.3. Topology optimization 3.5. Conclusion and outlook 3.6. References Chapter 4. Understanding Uncertainty Maurice LEMAIRE 4.1. Introduction 4.2. Uncertainty and uncertainties 4.3. Design and uncertainty 4.3.1. Decision modules 4.3.2. Designing in uncertain 4.4. Knowledge entity 4.4.1. Structure of a knowledge entity

Summary of Volume 3

4.5. Robust and reliable engineering 4.5.1. Definitions 4.5.2. Robustness 4.5.3. Reliability 4.5.4. Optimization 4.5.5. Reliable and robust optimization 4.6. Conclusion 4.7. References Chapter 5. New Approach to the Reliability Verification of Aerospace Structures Giora MAYMON 5.1. Introduction 5.2. Factor of safety and probability of failure 5.3. Reliability verification of aerospace structural systems 5.3.1. Reliability demonstration is integrated into the design process 5.3.2. Analysis of failure mechanism and failure modes 5.3.3. Modeling the structural behavior, verifying the model by tests 5.3.4. Design of structural development tests to surface failure modes 5.3.5. Design of development tests to find unpredicted failure modes 5.3.6. “Cleaning” failure mechanism and failure modes 5.3.7. Determination of required safety and confidence in models 5.3.8. Determination of the reliability by “orders of magnitude” 5.4. Summary 5.5. References Chapter 6. A Review of Interval Field Approaches for Uncertainty Quantification in Numerical Models Matthias FAES, Maurice I MHOLZ , Dirk VANDEPITTE and David M OENS 6.1. Introduction 6.2. Interval finite element analysis 6.3. Convex-set analysis 6.4. Interval field analysis 6.4.1. Explicit interval field formulation 6.4.2. Interval fields based on KL expansion 6.4.3. Interval fields based on convex descriptors 6.5. Conclusion 6.6. Acknowledgments 6.7. References

Modern Trends in Structural and Solid Mechanics

Chapter 7. Convex Polytopic Models for the Static Response of Structures with Uncertain-but-bounded Parameters Zhiping QIU and Nan JIANG 7.1. Introduction 7.2. Problem statements 7.3. Analysis and solution of the convex polytopic model for the static response of structures 7.4. Vertex solution theorem of the convex polytopic model for the static response of structures 7.5. Review of the vertex solution theorem of the interval model for the static response of structures 7.6. Numerical examples 7.6.1. Two-step bar 7.6.2. Ten-bar truss 7.6.3. Plane frame 7.7. Conclusion 7.8. Acknowledgments 7.9. References Chapter 8. On the Interval Frequency Response of Cracked Beams with Uncertain Damage Roberta SANTORO 8.1. Introduction 8.2. Crack modeling for damaged beams 8.2.1. Finite element crack model 8.2.2. Continuous crack model 8.3. Statement of the problem 8.3.1. Interval model for the uncertain crack depth 8.3.2. Governing equations of damaged beams 8.3.3. Finite element model versus continuous model 8.4. Interval frequency response of multi-cracked beams 8.4.1. Interval deflection function in the FE model 8.4.2. Interval deflection function in the continuous model 8.5. Numerical applications 8.6. Concluding remarks 8.7. Acknowledgments 8.8. References

Summary of Volume 3

Chapter 9. Quantum-Inspired Topology Optimization Xiaojun WANG, Bowen NI and Lei WANG 9.1. Introduction 9.2. General statements 9.2.1. Density-based continuum structural topology optimization formulation 9.2.2. Characteristics of quantum computing 9.3. Topology optimization design model based on quantum-inspired evolutionary algorithms 9.3.1. Classic procedure of topology optimization based on the SIMP method and optimality criteria 9.3.2. The fundamental theory of a quantum-inspired evolutionary algorithm – DCQGA 9.3.3. Implementation of the integral topology optimization framework 9.4. A quantum annealing operator to accelerate the calculation and jump out of local extremum 9.5. Numerical examples 9.5.1. Example of a short cantilever 9.5.2. Example of a wing rib 9.6. Conclusion 9.7. Acknowledgments 9.8. References Chapter 10. Time Delay Vibrations and Almost Sure Stability in Vehicle Dynamics Walter V. WEDIG 10.1. Introduction to road vehicle dynamics 10.2. Delay resonances of half-car models on road 10.3. Extensions to multi-body vehicles on a random road 10.4. Non-stationary road excitations applying sinusoidal models 10.5. Resonance reduction or induction by means of colored noise 10.6. Lyapunov exponents and rotation numbers in vehicle dynamics 10.7. Concluding remarks and main new results 10.8. References Chapter 11. Order Statistics Approach to Structural Optimization Considering Robustness and Confidence of Responses Makoto YAMAKAWA and Makoto OHSAKI 11.1. Introduction

Modern Trends in Structural and Solid Mechanics

11.2. Overview of order statistics 11.2.1. Definition of order statistics 11.2.2. Tolerance intervals and confidence intervals of quantiles 11.3. Robust design 11.3.1. Overview of the robust design problem 11.3.2. Worst-case-based method 11.3.3. Order statistics-based method 11.4. Numerical examples 11.4.1. Design response spectrum 11.4.2. Optimization of the building frame considering seismic responses 11.4.3. Multi-objective optimization considering robustness 11.5. Conclusion 11.6. References

Other titles from

in Mechanical Engineering and Solid Mechanics

2021 DAHOO Pierre Richard, POUGNET Philippe, EL HAMI Abdelkhalak Applications and Metrology at Nanometer Scale 1: Smart Materials, Electromagnetic Waves and Uncertainties (Reliability of Multiphysical Systems Set – Volume 9)

2020 LEDOUX Michel, EL HAMI ABDELKHALAK Heat Transfer 1: Conduction SALENÇON Jean Elastoplastic Modeling

2019 BAYLE Franck Reliability of Maintained Systems Subjected to Wear Failure Mechanisms: Theory and Applications (Reliability of Multiphysical Systems Set – Volume 8)

BEN KAHLA Rabeb, BARKAOUI Abdelwahed, MERZOUKI Tarek Finite Element Method and Medical Imaging Techniques in Bone Biomechanics (Mathematical and Mechanical Engineering Set – Volume 8) IONESCU Ioan R., QUEYREAU Sylvain, PICU Catalin R., SALMAN Oguz Umut Mechanics and Physics of Solids at Micro- and Nano-Scales LE VAN Anh, BOUZIDI Rabah Lagrangian Mechanics: An Advanced Analytical Approach MICHELITSCH Thomas, PÉREZ RIASCOS Alejandro, COLLET Bernard, NOWAKOWSKI Andrzej, NICOLLEAU Franck Fractional Dynamics on Networks and Lattices SALENÇON Jean Viscoelastic Modeling for Structural Analysis VÉNIZÉLOS Georges, EL HAMI Abdelkhalak Movement Equations 5: Dynamics of a Set of Solids (Non-deformable Solid Mechanics Set – Volume 5)

2018 BOREL Michel, VÉNIZÉLOS Georges Movement Equations 4: Equilibriums and Small Movements (Non-deformable Solid Mechanics Set – Volume 4) FROSSARD Etienne Granular Geomaterials Dissipative Mechanics: Theory and Applications in Civil Engineering RADI Bouchaib, EL HAMI Abdelkhalak Advanced Numerical Methods with Matlab® 1: Function Approximation and System Resolution (Mathematical and Mechanical Engineering SET – Volume 6) Advanced Numerical Methods with Matlab® 2: Resolution of Nonlinear, Differential and Partial Differential Equations (Mathematical and Mechanical Engineering SET – Volume 7)

SALENÇON Jean Virtual Work Approach to Mechanical Modeling

2017 BOREL Michel, VÉNIZÉLOS Georges Movement Equations 2: Mathematical and Methodological Supplements (Non-deformable Solid Mechanics Set – Volume 2) Movement Equations 3: Dynamics and Fundamental Principle (Non-deformable Solid Mechanics Set – Volume 3) BOUVET Christophe Mechanics of Aeronautical Solids, Materials and Structures Mechanics of Aeronautical Composite Materials BRANCHERIE Delphine, FEISSEL Pierre, BOUVIER Salima, IBRAHIMBEGOVIC Adnan From Microstructure Investigations to Multiscale Modeling: Bridging the Gap CHEBEL-MORELLO Brigitte, NICOD Jean-Marc, VARNIER Christophe From Prognostics and Health Systems Management to Predictive Maintenance 2: Knowledge, Traceability and Decision (Reliability of Multiphysical Systems Set – Volume 7) EL HAMI Abdelkhalak, RADI Bouchaib Dynamics of Large Structures and Inverse Problems (Mathematical and Mechanical Engineering Set – Volume 5) Fluid-Structure Interactions and Uncertainties: Ansys and Fluent Tools (Reliability of Multiphysical Systems Set – Volume 6) KHARMANDA Ghias, EL HAMI Abdelkhalak Biomechanics: Optimization, Uncertainties and Reliability (Reliability of Multiphysical Systems Set – Volume 5)

LEDOUX Michel, EL HAMI Abdelkhalak Compressible Flow Propulsion and Digital Approaches in Fluid Mechanics (Mathematical and Mechanical Engineering Set – Volume 4) Fluid Mechanics: Analytical Methods (Mathematical and Mechanical Engineering Set – Volume 3) MORI Yvon Mechanical Vibrations: Applications to Equipment

2016 BOREL Michel, VÉNIZÉLOS Georges Movement Equations 1: Location, Kinematics and Kinetics (Non-deformable Solid Mechanics Set – Volume 1) BOYARD Nicolas Heat Transfer in Polymer Composite Materials CARDON Alain, ITMI Mhamed New Autonomous Systems (Reliability of Multiphysical Systems Set – Volume 1) DAHOO Pierre Richard, POUGNET Philippe, EL HAMI Abdelkhalak Nanometer-scale Defect Detection Using Polarized Light (Reliability of Multiphysical Systems Set – Volume 2) DE SAXCÉ

Géry, VALLÉE Claude Galilean Mechanics and Thermodynamics of Continua DORMIEUX Luc, KONDO Djimédo Micromechanics of Fracture and Damage (Micromechanics Set – Volume 1) EL HAMI Abdelkhalak, RADI Bouchaib Stochastic Dynamics of Structures (Mathematical and Mechanical Engineering Set – Volume 2) GOURIVEAU Rafael, MEDJAHER Kamal, ZERHOUNI Noureddine From Prognostics and Health Systems Management to Predictive Maintenance 1: Monitoring and Prognostics (Reliability of Multiphysical Systems Set – Volume 4)

KHARMANDA Ghias, EL HAMI Abdelkhalak Reliability in Biomechanics (Reliability of Multiphysical Systems Set –Volume 3) MOLIMARD Jérôme Experimental Mechanics of Solids and Structures RADI Bouchaib, EL HAMI Abdelkhalak Material Forming Processes: Simulation, Drawing, Hydroforming and Additive Manufacturing (Mathematical and Mechanical Engineering Set – Volume 1)

2015 KARLIČIĆ Danilo, MURMU Tony, ADHIKARI Sondipon, MCCARTHY Michael Non-local Structural Mechanics SAB Karam, LEBÉE Arthur Homogenization of Heterogeneous Thin and Thick Plates

2014 ATANACKOVIC M. Teodor, PILIPOVIC Stevan, STANKOVIC Bogoljub, ZORICA Dusan Fractional Calculus with Applications in Mechanics: Vibrations and Diffusion Processes Fractional Calculus with Applications in Mechanics: Wave Propagation, Impact and Variational Principles CIBLAC Thierry, MOREL Jean-Claude Sustainable Masonry: Stability and Behavior of Structures ILANKO Sinniah, MONTERRUBIO Luis E., MOCHIDA Yusuke The Rayleigh−Ritz Method for Structural Analysis

LALANNE Christian Mechanical Vibration and Shock Analysis – 5-volume series – 3rd edition Sinusoidal Vibration – Volume 1 Mechanical Shock – Volume 2 Random Vibration – Volume 3 Fatigue Damage – Volume 4 Specification Development – Volume 5 LEMAIRE Maurice Uncertainty and Mechanics

2013 ADHIKARI Sondipon Structural Dynamic Analysis with Generalized Damping Models: Analysis ADHIKARI Sondipon Structural Dynamic Analysis with Generalized Damping Models: Identification BAILLY Patrice Materials and Structures under Shock and Impact BASTIEN Jérôme, BERNARDIN Frédéric, LAMARQUE Claude-Henri Non-smooth Deterministic or Stochastic Discrete Dynamical Systems: Applications to Models with Friction or Impact EL HAMI Abdelkhalak, RADI Bouchaib Uncertainty and Optimization in Structural Mechanics KIRILLOV Oleg N., PELINOVSKY Dmitry E. Nonlinear Physical Systems: Spectral Analysis, Stability and Bifurcations LUONGO Angelo, ZULLI Daniele Mathematical Models of Beams and Cables SALENÇON Jean Yield Design

2012 DAVIM J. Paulo Mechanical Engineering Education DUPEUX Michel, BRACCINI Muriel Mechanics of Solid Interfaces ELISHAKOFF Isaac et al. Carbon Nanotubes and Nanosensors: Vibration, Buckling and Ballistic Impact GRÉDIAC Michel, HILD François Full-Field Measurements and Identification in Solid Mechanics GROUS Ammar Fracture Mechanics – 3-volume series Analysis of Reliability and Quality Control – Volume 1 Applied Reliability – Volume 2 Applied Quality Control – Volume 3 RECHO Naman Fracture Mechanics and Crack Growth

2011 KRYSINSKI Tomasz, MALBURET François Mechanical Instability SOUSTELLE Michel An Introduction to Chemical Kinetics

2010 BREITKOPF Piotr, FILOMENO COELHO Rajan Multidisciplinary Design Optimization in Computational Mechanics DAVIM J. Paulo Biotribolgy

PAULTRE Patrick Dynamics of Structures SOUSTELLE Michel Handbook of Heterogenous Kinetics

2009 BERLIOZ Alain, TROMPETTE Philippe Solid Mechanics using the Finite Element Method LEMAIRE Maurice Structural Reliability

2007 GIRARD Alain, ROY Nicolas Structural Dynamics in Industry GUINEBRETIÈRE René X-ray Diffraction by Polycrystalline Materials KRYSINSKI Tomasz, MALBURET François Mechanical Vibrations KUNDU Tribikram Advanced Ultrasonic Methods for Material and Structure Inspection SIH George C. et al. Particle and Continuum Aspects of Mesomechanics